content
stringlengths
1
15.9M
\section{Introduction} The launch of fifth-generation (5G) wireless communication networks leads to an explosive increment of internet of things (IoT) devices, which will generate massive data that need to be processed timely either through local or remote computing facilities. Limited Bandwidth and limited device computation capability have gradually become a bottleneck to meet delay requirement for the massive amount of data. Meanwhile, the research on sixth-generation (6G) wireless networks has taken off and has been attracting ever increasing attention from academia and industry \cite{6g}. Future wireless networks aim for realizing ultra-high spectrum efficiency (SE) and energy efficiency (EE), ultra-dense user connectivity, and ultra low latency. EE is critically important in future wireless networks since massive energy constrained IoT devices need to be reliably connected to support smart city, smart factory, etc. Moreover, it is imperative to address the finite computation capacity issue of IoT devices. Mobile edge computing (MEC) is envisioned as a promising paradigm by exploiting the edge servers that are deployed close to the end-users. In wireless networks, edge nodes, such as base stations and edge routers, can be equipped with high computing and storage capabilities. MEC enables users to offload their tasks to edge servers for processing. It is important to have efficient offloading in MEC as massive local devices may need to offload tasks to edge servers for computing. Towards that end, NOMA can be applied to further enhance EE and SE of MEC networks. By exploiting superposition coding at transmitter and successive interference cancellation (SIC) at receiver, NOMA allows multiple users to share the same bandwidth simultaneously in either power domain or code domain to improve spectral usage \cite{hj1}. Extensive works have formulated EE optimization frameworks and designed energy-efficient resource allocation schemes in NOMA enabled MEC networks \cite{ee1}-\cite{ee3}. In order to maximize EE under the delay constraints in MEC, a joint radio and computation resource allocation problem was formulated in \cite{ee1}, which takes both intra and inter cell interference into consideration. In \cite{ee2}, the authors proposed a scheme to maximize the system computation EE of a wireless power transfer enabled NOMA based MEC network by jointly optimizing the computing frequencies, execution time, and transmit power. In \cite{ee3}, the task offloading and resource allocation problem in a NOMA assisted MEC network with security and EE was investigated. A dynamic task offloading and resource allocation algorithm was proposed based on Lyapunov optimization theory. However, complex wireless environment and devices' limited transmit power can render it very challenging for the NOMA based MEC network to achieve high offloading rate and high EE. IRS recently has emerged as a promising technique to tackle wireless capacity and energy issues. IRS consists of a large number of low-cost passive reflecting elements with the adjustable phase shifts \cite{irs1}. By properly adjusting the phase shifts of the elements, the reflected signals from various paths can be combined coherently to enhance the link achievable rate at the MEC receiver \cite{irsqun}. Through this, IRS can compensate for the pathloss and fading and allow users to use a relatively lower power to achieve a higher data rate. Since IRS does not employ any transmit radio frequency (RF) chains, energy consumption solely comes from reflective element phase adjustment and is usually very low. It becomes a very promising technology to improve EE in future wireless networks. The following studies have investigated the performance of using IRS with NOMA. In \cite{nomairs1}, an IRS-assisted uplink NOMA system was considered to maximize the sum rate of all the users under the individual power constraint. The considered problem requires a joint power control at the users and beamforming design at the IRS, and an SDR-based solution has been developed. In \cite{nomairs2}, the problem of joint user association, subchannel assignment, power allocation, phase shifts design, and decoding order determination was formulated for maximizing the sum-rate for an IRS-assisted NOMA network. In \cite{nomairs3}, an EE algorithm was proposed to yield a tradeoff between the rate maximization and power minimization for an IRS-assisted NOMA network. The authors aimed to maximize the system EE by jointly optimizing the transmit beamforming and the reflecting beamforming. It was shown that NOMA can improve EE compared to OMA. Furthermore, recently application of IRS into NOMA-based MEC networks has been studied. In \cite{nomairs4}, the authors investigated an IRS-aided MEC system with NOMA. By jointly optimizing the passive phase shifters, the size of transmission data, transmission rate, power, and time, as well as the decoding order, they aimed to minimize the sum energy consumption. A block coordinate descent method was developed to alternately optimize two separated subproblems. In \cite{nomairs5}, an IRS-aided MEC system was considered and a flexible time-sharing NOMA scheme was proposed to allow users to divide their data into two parts that are transmitted via NOMA and TDMA respectively. By designing the IRS passive reflection and users’ computation-offloading scheduling, the delay was minimized. However, neither \cite{nomairs4} nor \cite{nomairs5} considered the EE performance of IRS-assisted MEC networks with NOMA, which is very important for system design to obtain the optimal trade-off between achievable rate and consumed power. Motivated by the above-mentioned observations, in this paper, the EE maximization problem is studied in an IRS-assisted MEC network with NOMA. To the authors' best knowledge, this is the first work that focuses on EE performance for applying both NOMA and IRS in the MEC network. The major contributions of this paper are summarized as follows We investigate the joint design of the receiver beamforming, offloading power, phase shift matrix, and local computing frequency to maximize the EE in an IRS-assisted MEC network with NOMA. The problem is challenging to solve due to its non-convexity fractional objective function and coupling of the beamforming vector with the IRS phase shift matrix. An alternating optimization algorithm is proposed to solve the non-convex fractional problem by using semidefinite programming relaxation (SDR). The simulation results show that the proposed method can achieve the highest EE among all the benchmarks The rest of the paper is organized as follows. The system model is provided in Section II. The EE maximization problem and its solution are presented in Section III. Simulation results are given in Section IV. The paper is concluded in Section V. \textit{Notation:} $\mathbb{C}^{M\times N}$ denotes the $M \times N $ complex-valued matrices space. $\mathcal{CN}(\mu, \sigma^2)$ denotes the distribution of complex Gaussian random variable with mean $\mu$ and variance $\sigma^2$. For a square matrix $\mathbf{X}$, the trace of $\mathbf{X}$ is denoted as $\text{Tr}(\mathbf{X})$ and rank($\mathbf{X}$) denotes the rank of matrix $\mathbf{X}$. $\angle(x)$ denotes the phase of complex number $x$. Matrices and vectors are denoted by boldface capital letters and boldface lower case letters. \section{System Model As shown in Fig. \ref{model}, an IRS-assisted MEC system is considered. There are $K$ single-antenna user equipments (UEs) in the system, which can do both local computing and data offloading. The access point (AP) with an MEC server is equipped with $N$ antennas and the IRS has $M$ reflecting elements. \begin{figure}[h] \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \centering \includegraphics[width=3.0in]{sm0330.eps} \caption{An IRS-aided MEC system with NOMA.\label{model}} \end{figure} \subsection{Offloading Model} The baseband equivalent channel from UE $k$ to IRS, IRS to AP, and UE $k$ to AP are denoted as ${\bf{h}}_{I,U,k} \in \mathbb{C}^{1 \times M}$, ${\bf{H}}_{B,I,k} \in \mathbb{C}^{M \times N}$, and ${\bf{h}}_{B,U,k} \in \mathbb{C}^{1 \times N}$, respectively. In this paper, IRS adjusts its elements to maximize the combined incident signal from each UE to the AP. The diagonal phase-shift matrix can be denoted as $\mathbf{\Theta} = \text{diag}(\exp({j\theta_{1}}), \exp({j\theta_{2}}), $ $\cdots, \exp({j\theta_{M}}))$, wherein its main diagonal, $\theta_{m}\in [0,2\pi)$, denotes the phase shift on the combined incident signal by its $m$th element, $m=1,2,...,M$ \cite{cwh3}. The transmitted signal from UE $k$ is given as $\sqrt{p_k} s_k$, where $\sqrt{p_k}$ denotes the transmit power and $s_k$ denotes the independent information. $\mathbf{m}_{B,k}\in \mathbb{C}^{N \times 1}$ denotes the receive beam vectors with unit norm, i.e., $\left\| \mathbf{m}_{B,k} \right\|^2=1$ \cite{nomairs1}. Therefore, the signal received at AP can be given as \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \mathbf{y}_{B,U}=\sum_{k=1}^K({\bf{h}}_{B,U,k}^H + {\bf{h}}_{I,U,k}^H\Theta {{\bf{H}}_{B,I,k}}){{\bf{m}}_{B,k}}\sqrt {{p_k}} {s_k} + n_{B,U,k}, \end{equation} where $n_{B,U,k} \sim \mathcal{CN}(0,\sigma^2)$ is the complex additive white Gaussian noise (AWGN) \cite{nomairs3}, \cite{chanel1}. NOMA is used to improve SE and mitigate the interference between different UEs. By exploiting the SIC techniques, the received signal at AP is sequentially decoded and the UE with the best channel conditions is firstly decoded. The channel of each UE includes a direct link and a reflect link. Since the reflect link depends on the unknown parameters $\bf{\Theta}$, the effective channels cannot be used to order the users at the receiver side. Similar to \cite{nomairs1}, we simply remove unknown reflect matrix by considering it as an identity matrix $\mathbf{I}$. UEs are then sorted based on this channel gain $|(\mathbf{h}_{B,U,k}^{H}+\mathbf{h}_{I,U,k}^{H}\mathbf{I}\mathbf{H}_{B,I,k})|$. Without loss of generality, we assume that UEs are sorted in an increasing order, i.e., $|(\mathbf{h}_{B,U,1}^{H}+\mathbf{h}_{I,U,1}^{H}\mathbf{I}\mathbf{H}_{B,I,1})|\le |(\mathbf{h}_{B,U,2}^{H}+\mathbf{h}_{I,U,2}^{H}\mathbf{I}\mathbf{H}_{B,I,2})|\le \cdots \le|(\mathbf{h}_{B,U,K}^{H}+\mathbf{h}_{I,U,K}^{H}\mathbf{I}\mathbf{H}_{B,I,K})|$. When decoding the signal for UE $k$, the signals from $i=1,2,\cdots, k-1$ are treated as interference. Thus, the signal to interference plus noise ratio (SINR) for UE $k$ is expressed as \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \gamma_{B,k} = \frac{p_k|(\mathbf{h}_{B,U,k}^{H}+\mathbf{h}_{I,U,k}^{H}\mathbf{\Theta}\mathbf{H}_{B,I,k})\mathbf{m}_{B,k}|^2}{\sum_{i=1}^{k-1}p_i|(\mathbf{h}_{B,U,i}^{H}+\mathbf{h}_{I,U,i}^{H}\mathbf{\Theta}\mathbf{H}_{B,I,i})\mathbf{m}_{B,i}|^2+\sigma^2}. \end{equation} The achievable offloading rate is \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} R^{off}_{k}=B\log_2(1+\gamma_{B,k}). \end{equation} \subsection{Local Processing Model} Let $C_{k} $ be the number of computation cycles required to process one bit of data for UE $k$ locally. UE can compute and transmit simultaneously. Let $f_{k}$ denote the computing frequency of the processor (cycles/second) \cite{hj1}. Therefore, the local computing rate can be given a \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} R_{k}^{loc}=\frac{f_{k}}{C_{k}}. \end{equation} The power consumption of local computing is modeled as a function of processor speed $f_{k}$. It can be given as $p_{k}^{loc}= \epsilon f_{k}^3$, where $\epsilon$ is effective capacitance coefficient of processor chip. \subsection{Energy Efficiency} The energy consumed by each UE consists of transmit power, local computing power, and circuit power consumption. Thus, the total power consumed by each UE is given as \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} P^{tot}_k= p_k+\epsilon f_{k}^3+P^{cn}_k, \end{equation} where $P^{cn}_k$ denotes the constant circuit power consumed for signal processing and it is assumed to be the same for all UEs. The total achievable rate for each UE is \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} R_k^{tot}=R_k^{off}+R_k^{loc}. \end{equation} According to \cite{cwh3}, EE is defined as \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \eta =\frac{\sum_{k=1}^{K}R_k^{tot}}{\sum_{k=1}^{K}P_k^{tot}}. \end{equation} In order to maximize the EE, the local CPU frequency, offloading power, decoding vectors, and the phase shift matrix need to be jointly optimized. \section{Resource Optimization} In this section, the EE maximization problem is studied by jointly optimizing the local CPU frequency, offloading power, decoding vectors, and phase shift matrix. An alternating algorithm is further proposed to tackle the formulated problem \vspace{-0.03in} \subsection{Problem Formulation} The EE maximization problem is formulated as \begin{subequations} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \label{P0} \begin{alignat}{5} \textbf{P}{_1:}~ &\max_{p_k,f_k,\mathbf{m}_{B,k},\mathbf{ \Theta}}~\eta\nonumber\\ s.t.~~&P_k^{tot} \le P^{th}_k,\\ &R_k^{tot} \ge R_{th},\\ &|\exp({j\theta_{m}})|=1,\\ &\left\| \mathbf{m}_{B,k} \right\|^2=1, \end{alignat} \end{subequations} where $R_{th}$ is the minimum required rate threshold. $P_k^{th}$ is the maximum available power of each UE. It is evident that problem $\textbf{P}_1$ is non-convex due to the fractional structure of the objective function and the non-convex constraints. In order to tackle it, an alternating algorithm is proposed. By introducing ${\mathbf{w}^H}=[w_1, w_2, \cdots, w_M]$, one has $\mathbf{h}_{I,U,k}^{H}\mathbf{\Theta}\mathbf{H}_{B,I,k}={\mathbf{w}^H}\mathbf{H}{_{B,k}}$, where $w_m=\exp({j\theta_{m}})$, $\mathbf{H}_{B,k}=\text{diag}(\mathbf{h}_{I,U,k}^H)\mathbf{H}_{B,I,k}$. Thus, the SINR of UE $k$ is given as $\gamma_{B,k}=\frac{a_0p_k|\overline{\mathbf{w}}^H\overline{\mathbf{H}}_{B,k}\mathbf{m}_{B,k}|^2}{a_0\sum_{i=1}^{k-1}p_i|\overline{\mathbf{w}}^H\overline{\mathbf{H}}_{B,i}\mathbf{m}_{B,i}|^2+1}$, where $a_0=1/\sigma^2$, $\overline{\mathbf{H}}_{B,k}=\left[ {\begin{array}{*{20}{c}} {{\mathbf{H}_{B,k}}} \\ {{\mathbf{h}_{B,U,k}}} \\ \end{array}} \right]$, $\overline{\mathbf{w}}^H=\exp({j\overline{w}})[\mathbf{w}^H,1]$, and $\overline{w}$ is an arbitrary phase rotation. The objective of the optimization problem can be transformed into \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \frac{\frac{B}{\ln2}(\ln(1+\sum_{k=1}^K{a_0p_k|\overline{\mathbf{w}}^H\overline{\mathbf{H}}_{B,k}\mathbf{m}_{B,k}|^2}))+\sum_{k=1}^KR_k^{loc}}{\sum_{k=1}^KP_k^{tot}}. \label{10} \end{equation} To tackle the complexity introduced by the logarithmic function of $R_k^{off}$ in (\ref{P0}\text{b}), Lemma 1 is introduced. First, we have \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{split} R_k^{off}&=\frac{B}{\ln2}[\ln(a_0\sum_{i=1}^{k}p_i|\overline{\mathbf{w}}^H\overline{\mathbf{H}}_{B,i}\mathbf{m}_{B,i}|^2+1)\\ &-\ln({a_0\sum_{i=1}^{k-1}p_i|\overline{\mathbf{w}}^H\overline{\mathbf{H}}_{B,i}\mathbf{m}_{B,i}|^2+1})]. \end{split} \end{equation} $\mathbf{Lemma~1}$: By introducing the function $\phi(t)=-tx+\ln t+1$ for any $x > 0$, one has \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} -\ln x=\max_{t>0}\phi(t). \end{equation} The optimal solution can be achieved at $t=1/x$. By setting $x={a_0}\sum\limits_{i = 1}^{k - 1} {{p_i}} |{\overline {\bf{w}} ^H}{\overline {\bf{H}} _{B,i}}{{\bf{m}}_{B,i}}{|^2} + 1$, and $t=t_{B,k}$, one has \begin{equation} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{split} R_k^{off}&=\frac{B}{\ln2}\max_{t_{B,k}>0}\phi_{B,k}(p_k,f_k,\mathbf{m}_{B,k},\overline{\mathbf{w}},t_{B,k})\\ &=\frac{B}{\ln2}[\ln(a_0\sum_{i=1}^{k}p_i|\overline{\mathbf{w}}^H\overline{\mathbf{H}}_{B,i}\mathbf{m}_{B,i}|^2+1)+ \ln ({t_{B,k}}) + 1\\ & - {t_{B,k}}({a_0}\sum\limits_{i = 1}^{k - 1} {{p_i}} |{\overline {\bf{w}} ^H}{\overline {\bf{H}}_{B,i}}{{\bf{m}}_{B,i}}{|^2} + 1) ]. \end{split} \end{equation} By further introducing a variable $\eta_1$ to deal the fractional structure of (\ref{10}), $\textbf{P}{_{1}}$ can be transformed into \begin{subequations} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{alignat}{5} \label{p3} \begin{split} \textbf{P}_{2}:~ &\max_{p_k,f_k,\mathbf{m}_{B,k},t_{B,k},\overline{\mathbf{w}}}~ [\sum_{k=1}^KR_k^{loc}+\frac{B}{\ln2}(\ln(1+\\ &\sum_{k=1}^K{a_0p_k|\overline{\mathbf{w}}^H\overline{\mathbf{H}}_{B,k}\mathbf{m}_{B,k}|^2}))]-\eta_{1}{\sum_{k=1}^KP_k^{tot}} \end{split}\nonumber\\ s.t.~~& (\ref{P0}\text{a}), (\ref{P0}\text{c}),(\ref{P0}\text{d}),\nonumber\\\ \begin{split} &\frac{B}{\ln2}\phi_{B,k}(p_k,f_k,\mathbf{m}_{B,k},\overline{\mathbf{w}},t_{B,k})+R_k^{loc} \ge R_{th}. \end{split} \end{alignat} \end{subequations} $\textbf{P}_{2}$ is still non-convex due to the coupling of variables. An alternating algorithm is proposed. To be specific, $p_k$ and $f_k$ are first optimized with a given $\mathbf{m}_{B,k}$, and $\overline{\mathbf{w}}$. $\mathbf{m}_{B,k}$ can then be optimized with the obtained $p_k,f_k$, and $\overline{\mathbf{w}}$. Further $\overline{\mathbf{w}}$ can be optimized with the obtained $p_k,f_k$, and $\mathbf{m}_{B,k}$. This process iteratively continues until convergence. \subsection{CPU Frequency and Offloading Power Optimization} With the given $\mathbf{m}_{B,k}$ and $\overline{\mathbf{w}}$, let $A_{B,k}=a_0|\overline{\mathbf{w}}^H\overline{\mathbf{H}}_{B,k}\mathbf{m}_{B,k}|^2$, the problem can be transformed into \begin{subequations} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{alignat}{5} \begin{split} \textbf{P}{_{3.1}:}~ &\max_{p_k,f_k}~\frac{B}{\ln2}(\ln(\sum_{k=1}^{K}p_kA_{B,k}+1))\\ &+\sum_{k=1}^{K}\frac{f_k}{C_k}-\eta_1 \sum_{k=1}^K(\zeta p_k+\epsilon f_{k}^3+P^{cn}_k) \end{split}\nonumber\\ s.t.~~& p_k+\epsilon f_{k}^3+P^{cn}_k \le P_k^{th},\\ \begin{split} &\frac{B}{\ln2}\phi_{B,k}(p_k,f_k)+\frac{f_k}{C_k} \ge R_{th}. \end{split} \end{alignat} \end{subequations} Problem $\textbf{P}_{3.1}$ is convex with respect to $f_k$ and $p_k$, therefore, it can be solved by using a standard convex optimization tool. \subsection{Optimizing the Receiving Beamforming} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} In this section, we solve the problem $\textbf{P}_{2}$ to achieve the receive beamforming vector $\mathbf{m}_{B,k}$ for a given $\overline{\mathbf{w}}$, $p_k$, and $f_k$. Let $\mathbf{\overline{h}}_{B,k}^H=\mathbf{\overline{w}}^H\mathbf{\overline{H}}_{B,k}$, $\textbf{P}_{2}$ can be transformed into \begin{subequations} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{alignat}{5} \label{p3.2} \begin{split} \textbf{P}_{3.2}:~ &\max_{\mathbf{m}_{B,k}}~\frac{\ln2}{B}\ln(a_0\sum_{k=1}^{K}p_k|\mathbf{\overline{h}}_{B,k}^H\mathbf{m}_{B,k}|^2\\ &+1)+\sum_{k=1}^KR_k^{loc}-\eta_{1}\sum_{k=1}^KP_k^{tot} \end{split}\nonumber\\ \begin{split} s.t.~~&\frac{B}{\ln2}\ln(a_0\sum_{i=1}^{k}p_i|\mathbf{\overline{h}}_{B,k}^H\mathbf{m}_{B,i}|^2+1)+ \ln ({t_{B,k}}) + 1\\ & - {t_{B,k}}({a_0}\sum\limits_{i = 1}^{k - 1} {{p_i}} |{\mathbf{\overline{h}}_{B,i}^H{{\bf{m}}_{B,i}}}{|^2} + 1) +R_k^{loc}\ge R_{th}, \end{split}\\ &\left\| \mathbf{m}_{B,k} \right\|^2=1. \end{alignat} \end{subequations} Let $|\overline{\mathbf{h}}_{B,k}^H\mathbf{m}_{B,k}|^2 = \rm{Tr}(\tilde{\mathbf{H}}_{B,k}\mathbf{m}_{B,k}\mathbf{m}_{B,k}^H)$. By defining $\tilde{\mathbf{H}}_{B,k}=\mathbf{\overline{h}}_{B,k}\mathbf{\overline{h}}_{B,k}^H$, $\mathbf{M}_{B,k} = \mathbf{m}_{B,k}\mathbf{m}_{B,k}^H$, one has $\mathbf{M}_{B,k}\succeq0$ and $\text{rank}(\mathbf{{M}}_{B,k})=1$. The rank-$1$ constraint makes the problem difficult to solve. Thus, we apply the SDR method to relax the constraints \cite{irsqun}. $\textbf{P}_{3.2}$ is then expressed as \begin{subequations} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{alignat}{5} \label{p3.3} \begin{split} \textbf{P}_{3.3}:~ &\max_{\mathbf{M}_{B,k}}~[\frac{B}{\ln2}\ln(a_0\sum_{k=1}^{K}p_k\rm{Tr}(\tilde{\mathbf{H}}_{B,k}{\mathbf{M}_{B,k}})\\ &+1) +\sum_{k=1}^KR_k^{loc}]-\eta_{1}\sum_{k=1}^KP_k^{tot} \end{split}\nonumber\\ \begin{split} s.t.~~&\frac{B}{\ln2}\ln(a_0\sum_{i=1}^{k}p_i\rm{Tr}(\tilde{\mathbf{H}}_{B,i}{\mathbf{M}_{B,i}})+1)+ \ln ({t_{B,k}}) + 1\\ & - {t_{B,k}}({a_0}\sum\limits_{i = 1}^{k - 1} {{p_i}} \rm{Tr}(\tilde{\mathbf{H}}_{B,i}{\mathbf{M}_{B,i}}) + 1) +R_k^{loc}\ge R_{th}, \end{split}\\ &\rm{Tr}({\mathbf{M}_{B,k}})=1. \end{alignat} \end{subequations} $\textbf{P}_{3.3}$ is convex and can be solved by using a standard convex optimization tool \cite{wqqirs}. After $\mathbf{M}_{B,k}$ is obtained, if $\text{rank}(\mathbf{M}_{B,k})=1$, $\mathbf{m}_{B,k}$ can be obtained from $\mathbf{M}_{B,k}=\mathbf{m}_{B,k}\mathbf{m}_{B,k}^H$ by performing the eigenvalue decomposition. Otherwise, the Gaussian randomization can be used for recovering $\mathbf{m}_{B,k}$ \cite{wqqirs}. \subsection{Optimizing the IRS Reflecting Shifts $\overline{\mathbf{w}}$} After obtaining the beamforming vectors $\mathbf{m}_{B,k}$, by setting $\mathbf{h}_{W,B,k}=\mathbf{\overline{H}}_{B,k}\mathbf{m}_{B,k}$, problem $\textbf{P}_{2}$ can be transformed into \begin{subequations} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{alignat}{5} \begin{split} \textbf{P}{_{3.4}:}~ &\max_{\overline{\mathbf{w}}}~\frac{B}{\ln2}\ln(a_0\sum_{k=1}^{K}p_k(\overline{\mathbf{w}}^H\mathbf{h}_{W,B,k})+1)\\ & +\sum_{k=1}^K\frac{f_k}{C_k} -\eta_{1}\sum_{k=1}^KP_k^{tot} \end{split}\nonumber\\ s.t.~~& |w_{m}|=1,~m=1,2,...,M,\\ \begin{split} &\ln(a_0\sum_{i=1}^{k}p_i(\overline{\mathbf{w}}^H\mathbf{h}_{W,B,i})+1)+ \ln ({t_{B,k}}) + 1 \\ &- {t_{B,k}}({a_0}\sum\limits_{i = 1}^{k - 1} {{p_i}} (\overline{\mathbf{w}}^H\mathbf{h}_{W,B,i}) + 1) +\frac{f_k}{C_k} \ge R_{th}. \end{split} \end{alignat} \end{subequations} Similar to the previous section, let $\mathbf{W}=\overline{\mathbf{w}}\overline{\mathbf{w}}^H$, $\mathbf{H}_{W,B,k}=\mathbf{h}_{W,B,k}\mathbf{h}_{W,B,k}^H$. By applying the SDR method, we have \begin{subequations} \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \begin{alignat}{5} \begin{split} \textbf{P}{_{3.5}:}~ &\max_{\mathbf{W}}~\frac{B}{\ln2}\ln(a_0\sum_{k=1}^{K}p_k\text{Tr}(\mathbf{H}_{W,B,k}\mathbf{W})+1)\\ & +\sum_{k=1}^K\frac{f_k}{C_k} -\eta_{1}\sum_{k=1}^KP_k^{tot} \end{split}\nonumber\\ s.t.~~& \mathbf{W} \succeq 0, \mathbf{W}_{mm}=1,~m=1,2,...,M,\\ \begin{split} &\ln(a_0\sum_{i=1}^{k}p_i\text{Tr}(\mathbf{H}_{W,B,k}\mathbf{W})+1)+ \ln ({t_{B,k}}) + 1 \\ &- {t_{B,k}}({a_0}\sum\limits_{i = 1}^{k - 1} {{p_i}} \text{Tr}(\mathbf{H}_{W,B,k}\mathbf{W}) + 1)+\frac{f_k}{C_k} \ge R_{th}. \end{split} \end{alignat} \end{subequations} The problem $\textbf{P}_{3.5}$ is a convex problem and can be solved by using a standard convex optimization tool. After obtaining $\mathbf{W}$, $\overline{\mathbf{w}}$ can be given by eigenvalue decomposition if $\text{rank}({\mathbf{W}})=1$; otherwise, the Gaussian randomization can be used for recovering the approximate $\overline{\mathbf{w}}$ \cite{wqqirs}. The reflection coefficients can be given by $w_m=\angle(\frac{\overline{{w}}_m}{\overline{{w}}_{M+1}}),~m=1,2,..,M$. The overall optimization algorithm is summarized in Algorithm 1, where $\delta$ is the threshold and $T$ is the maximum number of iterations \begin{table}[htbp] \setlength{\abovedisplayskip}{3pt} \setlength{\belowdisplayskip}{3pt} \centering \begin{center} \begin{tabular}{lcl} \\\toprule { $\textbf{Algorithm 1}$: Alternating Algorithm for Solving $\textbf{P}_{1}$}\\ \midrule \ 1) \textbf{Input settings:}\\ \ \ \ \ \ \ \ $\delta$, $R_{th},P_k^{th} >0$, and $T$.\\ \ 2) \textbf{Initialization:}\\ \ \ \ \ \ \ \ $t_{B,k}(0)$, $\mathbf{\overline{w}}(0)$,$\mathbf{m}_{B,k}(0)$, and $\eta_1(0)$;\\ \ 3) \textbf{Optimization:}\\ \ \ \ \ \ \textbf{$\pmb{\unrhd} $ for \emph{$\tau_1$}=1:T }\\ \ \ \ \ \ \ \ \ \ \ solve $\textbf{P}_{3.1}$ with $\mathbf{\overline{w}}^{*}(\tau_1-1)$,$\mathbf{m}_{B,k}^*(\tau_1-1)$;\\ \ \ \ \ \ \ \ \ \ \ obtain the solution $p_k^*(\tau_1)$, $f_k^*(\tau_1)$;\\ \ \ \ \ \ \ \ \ \ \ solve $\textbf{P}_{3.3}$ with $p_k^*(\tau_1)$, $f_k^*(\tau_1)$, and $\mathbf{\overline{w}}^{*}(\tau_1-1)$;\\ \ \ \ \ \ \ \ \ \ \ obtain the solution $\mathbf{m}_{B,k}^*(\tau_1)$;\\ \ \ \ \ \ \ \ \ \ \ solve $\textbf{P}_{3.5}$ with $p_k^*(\tau_1)$, $f_k^*(\tau_1)$, and $\mathbf{m}_{B,k}^*(\tau_1)$;\\ \ \ \ \ \ \ \ \ \ \ obtain the solution $\mathbf{\overline{w}}^{*}(\tau_1)$;\\ \ \ \ \ \ \ \ \ \ \ calculate EE $\eta(\tau_1)$ and update $t_{B,k}(\tau_1)$ and $\eta_1(\tau_1)$;\\ \ \ \ \ \ \ \ \ \ \ \textbf{if} $| \frac{\eta(\tau_1)-\eta(\tau_1-1)}{\eta(\tau_1)}| \le \delta$; \\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ the optimal EE $\eta^*$ is obtained;\\ \ \ \ \ \ \ \ \ \ \ \textbf{end}\\ \ \ \ \ \ \textbf{$\pmb{\unrhd} $ end} \\ \ 4) \textbf{Output:}\\ \ \ \ \ \ \ \ $p_k^*$, $f_k^*$, $\mathbf{m}_{B,k}^*$, and $\mathbf{w}^{*}$ and EE $\eta^*$.\\ \bottomrule \end{tabular} \end{center} \end{table} \section{Simulation Results} In this section, simulation results are provided to evaluate the performance of the proposed algorithms. We consider a three-dimensional Cartesian coordinate system. The simulation settings are based on those used in \cite{irs1}, \cite{wqqirs}. We consider a $2$-UE case and it can be readily extend to multiple UE cases. The locations of the MEC, the IRS, UE1, and UE2 are set as $(5,0,20)$, $(0,50,2)$, $(5,75,5)$ and $(5,50,10)$, respectively \cite{wqqirs}. The channels are generated by $h_{i,j}=\sqrt{G_0d_{i,j}^{-c_{i,j}}}g_{i,j}$, where $G_0=-30$ dB is the path loss at the reference point. $d_{i,j}$, $c_{i,j}$ and $g_{i,j}$ denote the distance, path loss exponent, and fading between $i$ and $j$, respectively, where $i \in \{B, I\}$ and $j \in \{U,k\}$. The path loss exponents are set as $c_{B,U,k}=5$, $c_{B,I}=3.5$, and $c_{I,U,k}=2$. The bandwidth $B$ is set to $1$ Mhz. Other parameters are set as $\sigma^2= -105$ dBm, $P_k^{th}=31$ dBm, $P_k^{cn}=23$ dBm, $C_k=10^3$ cycles/bit, and $\epsilon=10^{-28}$. The proposed scheme is marked as `Efficiency-IRS'. We consider three other cases as benchmarks to compare with the proposed method. The first benchmark, marked as `OMA-IRS', uses FDMA with equally allocated bandwidth to all the users. The second benchmark, marked as `OnlyOff-IRS', has no local computing and all the tasks are offloaded. The third benchmark, marked as `Efficiency-NoIRS", aims to investigate the performance without IRS. \begin{figure}[h] \setlength{\abovecaptionskip}{-0.2cm} \setlength{\belowcaptionskip}{-1cm} \centering \includegraphics[width=3.0in]{etarth.eps} \caption{Energy efficiency versus the minimum rate threshold.\label{etarth}} \end{figure} Fig. \ref{etarth} shows EE versus the minimum rate threshold $R_{th}$. The EE achieved by the proposed method is the best among all the schemes. This indicates that the IRS assisted MEC with NOMA can help improve the system rate and achieve high EE. With the increase of $R_{th}$, all the curves are decreasing. The system has to consume excessive energy to increase the rate in order to meet the minimum rate constraint, which decreases EE. \begin{figure}[h] \setlength{\abovecaptionskip}{-0.2cm} \setlength{\belowcaptionskip}{-1cm} \centering \includegraphics[width=3.0in]{raterth.eps} \caption{Achievable rate versus the minimum rate threshold.\label{raterth}} \end{figure} A comparison of the system rate versus the rate threshold $R_{th}$ is presented in Fig. \ref{raterth}. All the curves increase with $R_{th}$ in order to meet the service requirement, which verifies the observation in Fig. \ref{etarth}. The system rates obtained by the proposed method and the `OnlyOff-IRS' method are higher than those of the other two methods, which indicates that combining IRS with NOMA can significantly help the system to achieve a higher rate. It is worth noting that even though the `OnlyOff-IRS' method can achieve the highest rate when $R_{th}$ is low, its efficiency is lower than the proposed method. This indicates that the overall efficiency performance degrades when there is no local computing. \begin{figure}[h] \setlength{\abovecaptionskip}{-0.2cm} \setlength{\belowcaptionskip}{-1cm} \centering \includegraphics[width=3.0in]{powerrth.eps} \caption{Power consumption versus the minimum rate threshold.\label{powerrth}} \end{figure} \begin{figure}[h] \setlength{\abovecaptionskip}{-0.2cm} \setlength{\belowcaptionskip}{-1cm} \centering \includegraphics[width=3.0in]{irsdis.eps} \caption{Energy efficiency versus the relative distance of UE-IRS.\label{etairs}} \end{figure} \begin{figure}[h] \setlength{\abovecaptionskip}{-0.2cm} \setlength{\belowcaptionskip}{-1cm} \centering \includegraphics[width=3.0in]{iters.eps} \caption{Convergence with Iteration.\label{iters}} \end{figure} Fig. \ref{powerrth} presents the power consumption versus $R_{th}$ for different methods. The results of all the methods in Fig. \ref{powerrth} are consistent with what are shown in Fig. \ref{etarth} and Fig. \ref{raterth}. It is worth noting that the power consumption by the proposed method is quite low. So UEs can use less energy to achieve a higher rate, which demonstrates the advantage of combining NOMA and IRS to MEC network in improving EE. Fig. \ref{etairs} shows EE versus the distance between UEs and IRS. The distance is the relative increased amount compared with UEs' original position. The curves for all the methods with IRS decrease with the increase of the distance except `Efficiency-NoIRS'. This is because the increase of the distance results in the increase of the path loss and the reduction of the power gain from the reflecting path through the IRS. Therefore, the achievable rate and EE are both decreased. It can also be seen that the `Efficiency-IRS' method still has the highest performance among all the methods, which validates the superiority of the proposed design. In Fig. \ref{iters}, the coverage of proposed methods based on different $R_{th}$ setting are investigated. It can be observed from Fig. \ref{iters} that only several iterations are needed for the proposed algorithms to converge, showing the computation efficiency of the proposed algorithm. \section{Conclusions} In this paper, an IRS-assisted MEC network with NOMA was considered. EE was maximized by jointly optimizing the offloading power, local computing frequency, beamforming vectors, and IRS phase shift matrix. An alternating algorithm was proposed to solve the challenging non-convex fractional optimization problems. The numerical results showed that our proposed method outperforms other benchmark schemes in terms of EE. It was proved that NOMA and IRS could help the MEC network to achieve a higher rate with a lower power. The convergence of the proposed algorithm was also verified.
\section{#1}} \renewcommand{\theequation}{\thesection.\arabic{equation}} \newcommand{\reff}[1]{(\ref{#1})} \newcommand{\vl}[2]{\rule[#1]{0.1mm}{#2}} \newcommand{\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}}{\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}} \newcommand{\!\stackrel{\textstyle{\hst{0ex} \atop \circ}}{\hst{0ex}}\!}{\!\stackrel{\textstyle{\hst{0ex} \atop \circ}}{\hst{0ex}}\!} \newcommand{\Gamma}{\Gamma} \newcommand{\Delta}{\Delta} \newcommand{\alpha}{\alpha} \newcommand{\beta}{\beta} \newcommand{\gamma}{\gamma} \newcommand{\delta}{\delta} \newcommand{\varepsilon}{\varepsilon} \newcommand{\varrho}{\varrho} \newcommand{\sigma}{\sigma} \newcommand{\varphi}{\varphi} \newcommand{\omega}{\omega} \newcommand{\Gamma^{\ast}}{\Gamma^{\ast}} \newcommand{\Co\Gas}{\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}\Gamma^{\ast}} \newcommand{\Co\Ga}{\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}\Gamma} \newcommand{\langle}{\langle} \newcommand{\rangle}{\rangle} \newcommand{\rightarrow}{\rightarrow} \newcommand{\longrightarrow}{\longrightarrow} \newcommand{\otimes}{\otimes} \begin{document} \begin{titlepage} \hspace*{\fill} \parbox{8em}{BONN--TH--95--16 \\ August 1995 } \vspace{18mm} \begin{center} {\LARGE\bf Universal $R$--matrices \\[-1.5ex] for finite Abelian groups --- \\[1ex] a new look at graded multilinear algebra} \\ \vspace{11mm} {\large M. Scheunert} \\ \vspace{5mm} Physikalisches Institut der Universit\"{a}t Bonn \\ Nu{\ss}allee 12, D--53115 Bonn, Germany \end{center} \vspace{20mm} \begin{abstract} \noindent The universal $R$--matrices and, dually, the coquasitriangular structures of the group Hopf algebra of a finite Abelian group (resp. of an arbitrary Abelian group) are determined. This is used to formulate graded multilinear algebra in terms of triangular or cotriangular Hopf algebras. For the convenience of the reader, in a separate section the definitions and basic properties of quasitriangular and coquasitriangular Hopf algebras are recalled. \end{abstract} \vspace{\fill} q-alg/9508016 \end{titlepage} \setcounter{page}{2} \section{Introduction} In the present work I am going to discuss the universal $R$--matrices [1] for finite Abelian groups, or, more precisely, the universal $R$--matrices for the group Hopf algebras of these groups [2]. To explain how I became interested in this problem I remind the reader of the graded multilinear algebra described in Ref. [3]. As shown in this reference, given an (arbitrary) Abelian group $\Gamma$ and a commutation factor $\sigma$ on $\Gamma$, the basic constructions of classical multilinear algebra can easily be generalized to the $\Gamma$--graded setting. In modern terms, this means that the $\Gamma$--graded vector spaces form a tensor category (depending on $\sigma$) with an internal hom, and this category is rigid if we consider only finite--dimensional vector spaces. It is this category where the so--called Lie colour algebras (also called $\sigma$ Lie algebras) live [4,\,5], and that is why I have investigated these structures. Since the present discussion is only meant to serve as a motivation, I do not want to explain the various category theoretical terms used above, but refer to Refs. [6,\,7,\,8], where more details are given. However, I would like to emphasize that the properties of the categories in question are closely related to the axioms for a triangular Hopf algebra [1]. In fact, given any triangular Hopf algebra $H$, the (left) $H$--modules form a category of this type, and conversely, under suitable assumptions, such a category is equivalent to the category of $H$--modules for some triangular Hopf algebra $H$ (this is the content of various so--called reconstruction theorems). In the simple case under consideration we do not want to invoke one of the reconstruction theorems in order to find the appropriate triangular Hopf algebra. Instead, we recall that, for a finite Abelian group $\Gamma$, a $\Gamma$--graded vector space is nothing but a vector space endowed with a representation of $\Gamma^{\ast}$, the character group of $\Gamma$ (see Section 2). Thus the group Hopf algebra $\Co\Gas$ of $\Gamma^{\ast}$ is the natural candidate for $H$, and we have to determine the triangular universal $R$--matrices for $\Co\Gas$. (Recall that, for a finite Abelian group $\Gamma$, the groups $\Gamma$ and $\Gamma^{\ast}$ are isomorphic.) Actually, we do more and determine all the universal $R$--matrices for $\Co\Gas$. From the category theoretical point of view this means the transition from (symmetric) tensor categories to quasitensor or braided monoidal categories. The outcome is completely satisfactory: There exists a simple bijection of the set of bicharacters on $\Gamma$ onto the set of universal $R$--matrices for $\Co\Gas$, and this bijection maps the commutation factors onto the triangular universal $R$--matrices. The proof given in Section 2 is elementary, it only uses some simple properties of finite Abelian groups and their characters. Nevertheless, it is worth--while to consider the problem from a dual point of view. It is easy to see that the group Hopf algebras $\Co\Ga$ and $\Co\Gas$ are dual to each other. Thus the $\Gamma$--graded vector spaces, i.e., the $\Co\Gas$--modules, can be identified with the $\Co\Ga$--comodules, and the universal $R$--matrices for $\Co\Gas$ correspond to coquasitriangular structures on $\Co\Ga$. Once this has been realized, the results summarized above follow immediately. Moreover, in the dual language, these results hold also for not necessarily finite Abelian groups. All this will be shown in Section 4, after some of the notions used there (and above) have been explained in the preparatory Section 3. The final Section 5 contains some concluding remarks and draws the reader's attention to related work by other authors. In closing I would like to mention some notations and conventions which will be used throughout this work. The base field will always be the field $\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}$ of complex numbers, and the multiplicative group of non--zero complex numbers will be denoted by $\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}_{\ast}$\,. Moreover, $\Gamma$ will be an Abelian group which, for convenience, will be written {\it multiplicatively}. A character of $\Gamma$ is a homomorphism of $\Gamma$ into $\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}_{\ast}$\,. The characters of $\Gamma$ form a (multiplicative) group $\Gamma^{\ast}$, and the canonical pairing of $\Gamma^{\ast}$ and $\Gamma$ will be denoted by a pointed bracket: \be \langle\,\;,\;\rangle : \Gamma^{\ast} \times \Gamma \longrightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}_{\ast} \end{equation} \be \langle\gamma',\gamma\rangle = \gamma'(\gamma) \;\;\mbox{for all}\;\; \gamma'\in\Gamma^{\ast}\,,\;\gamma\in\Gamma. \end{equation} Of course, if $\Gamma$ is finite, its characters take their values in the unit circle $U(1)$. \section{Universal $R$--matrices for the group \newline Hopf algebra of a finite Abelian group} \setcounter{equation}{0} {\it In the present section the group} $\Gamma$ {\it is assumed to be finite.} Let us first show that there exists a bijective correspondence between $\Gamma$--graded vector spaces and $\Gamma^{\ast}$--modules, i.e., vector spaces endowed with a representation of $\Gamma^{\ast}$. In fact, suppose $V$ is a $\Gamma^{\ast}$--module and let $\gamma'\cdot\,x$ denote the action of an element $\gamma'\in\Gamma^{\ast}$ on an element $x\in V$. Since $\Gamma^{\ast}$ is finite, any representation of $\Gamma^{\ast}$ is completely reducible, and since $\Gamma^{\ast}$ is Abelian, its irreducible representations are one--dimensional, hence given by its characters, i.e., by the elements of $\Gamma$. This implies that $V$ has the decomposition \be V = \bigoplus_{\gamma\in\Gamma}V_{\gamma}\:, \label{dec} \end{equation} where the subspaces $V_{\gamma}$ of $V$ are defined by \be V_{\gamma} = \{x \in V\,|\,\gamma'\cdot\,x = \langle\gamma',\gamma\rangle\,x \;\;\mbox{for all}\;\;\gamma'\in\Gamma^{\ast}\}\,. \end{equation} Conversely, suppose that $V$ is a $\Gamma$--graded vector space, i.e., suppose that a decomposition like \reff{dec} is given. Then the action of $\Gamma^{\ast}$ on $V_{\gamma}$ defined by \be \gamma'\cdot\,x = \langle\gamma',\gamma\rangle\,x\;\;\mbox{for all}\;\;\gamma'\in\Gamma^{\ast}\,,\; x\in V_{\gamma} \end{equation} makes each of the subspaces $V_{\gamma}$ of $V$ into a $\Gamma^{\ast}$--module, and $V$ can be considered as the direct sum of these modules. It is easy to see that these transitions from $\Gamma^{\ast}$--modules to $\Gamma$--graded vector spaces and vice versa are inverse to each other. Moreover, the homomorphisms of $\Gamma^{\ast}$--modules are exactly the homomorphisms of $\Gamma$--graded vector spaces, i.e., the linear mappings which are homogeneous of degree zero. Now let $\Co\Gas$ be the group algebra of $\Gamma^{\ast}$. Recall that $\Co\Gas$ is the associative algebra consisting of the formal linear combinations of elements of $\Gamma^{\ast}$, with the multiplication inherited from $\Gamma^{\ast}$, recall also that $\Gamma^{\ast}$--modules and $\Co\Gas$--modules are essentially the same thing. It is well--known [2] that $\Co\Gas$ is a Hopf algebra, where the coproduct $\Delta$, the counit $\varepsilon$, and the antipode $S$ are given (on the basis $\Gamma^{\ast}$ of $\Co\Gas$) by \begin{eqnarray} \Delta(\gamma') \!\! & = & \!\! \gamma'\otimes\gamma' \\ \varepsilon(\gamma') \!\! & = & \!\! 1 \\ S(\gamma') \!\! & = & \!\! \gamma'^{-1} \end{eqnarray} for all $\gamma'\in\Gamma^{\ast}$. Let us now proceed to the determination of the universal $R$--matrices for $\Co\Gas$. To find a good Ansatz, let $\sigma$ be a commutation factor on $\Gamma$ and let $V$ and $W$ be two $\Gamma$--graded vector spaces. Then, according to Ref. [3], the canonical isomorphism \be \psi_{V,W} : V \otimes W \longrightarrow W \otimes V \end{equation} is given by \be \psi_{V,W}(x \otimes y) = \sigma(\alpha,\beta)\,y \otimes x \label{col} \end{equation} for all $x \in V_{\alpha}$\,, $y \in W_{\beta}$\,; $\alpha,\beta \in \Gamma$. On the other hand, for a triangular Hopf algebra $H$ with universal $R$--matrix \be R = \sum_{i}R^1_i \otimes R^2_i \end{equation} (with $R^1_i, R^2_i \in H$) and for any two $H$--modules $V$ and $W$, the corresponding braiding isomorphism is fixed by \be \psi_{V,W}(x \otimes y) = \sum_{i}R^2_i\cdot y \otimes R^1_i\cdot x \label{braid} \end{equation} for all $x \in V$, $y \in W$, where, as before, the dot denotes the action of an element of $H$ on some element of an $H$--module. In our case, we have $H = \Co\Gas$, and a general element of $\Co\Gas \otimes \Co\Gas$ can be uniquely written in the form \be R = \sum_{\alpha',\,\beta'\in \Gamma^{\ast}}c_{\alpha',\,\beta'}\,\alpha'\otimes \beta' \; , \end{equation} with some coefficients $c_{\alpha',\,\beta'} \in \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}$\,. Inserting this expression into Eq. \reff{braid} and equating the right hand sides of Eqs. \reff{col} and \reff{braid} we obtain the condition \be \sigma(\alpha,\beta) = \sum_{\alpha',\,\beta'}c_{\alpha',\,\beta'}\, \langle\alpha',\alpha\rangle \langle\beta',\beta\rangle \label{cond} \end{equation} for all $\alpha,\beta \in \Gamma$. Using the well--known orthogonality and completeness relations for the characters, i.e., \be \sum_{\alpha \in \Gamma}\langle\alpha',\alpha\rangle\langle\beta',\alpha\rangle^{\ast} = n \delta_{\alpha',\,\beta'} \label{orth} \end{equation} \be \sum_{\alpha' \in \Gamma^{\ast}}\langle\alpha',\alpha\rangle\langle\alpha',\beta\rangle^{\ast} = n \delta_{\alpha,\,\beta}\;, \label{compl} \end{equation} where $\alpha',\beta' \in \Gamma^{\ast}$ and $\alpha,\beta \in \Gamma$ and where $n$ denotes the order of $\Gamma$, we can solve Eq. \reff{cond} for the coefficients $c_{\alpha',\,\beta'}$ and obtain the following Ansatz for the universal $R$--matrix correseponding to $\sigma$\,: \be R_{\sigma} = \frac{1}{n^2}\sum_{\alpha,\beta \in \Gamma \atop \alpha',\,\beta' \in \Gamma^{\ast}} \sigma(\alpha,\beta)\langle\alpha',\alpha\rangle^{\ast}\langle\beta',\beta\rangle^{\ast}\, \alpha' \otimes \beta'\,. \label{ans} \end{equation} It turns out that Eq. \reff{ans} gives a nice parametrization of the elements of $\Co\Gas \otimes \Co\Gas$. Let ${\cal F}(\Gamma \times \Gamma)$ be the algebra of all complex--valued functions on $\Gamma \times \Gamma$. Then we can prove the following proposition. \\ \noindent {\bf Proposition 2.1} \\ {\it 1) The mapping $\sigma \rightarrow R_{\sigma}$ of ${\cal F}(\Gamma \times \Gamma)$ into} $\Co\Gas \otimes \Co\Gas$ {\it given by Eq. \reff{ans} is an algebra isomorphism.} \noindent In the subsequent statements, $\sigma$ denotes an arbitrary element of ${\cal F}(\Gamma \times \Gamma)$. \\ {\it 2) We have \be (\Delta \otimes id)(R_{\sigma}) = (R_{\sigma})_{13}(R_{\sigma})_{23} \label{UR1} \end{equation} if and only if \be \sigma(\alpha\beta,\gamma) = \sigma(\alpha,\gamma)\sigma(\beta,\gamma)\;\;\mbox{\it for all}\;\; \alpha,\beta,\gamma \in \Gamma\,. \label{bich1} \end{equation} 3) We have \be (id \otimes \Delta)(R_{\sigma}) = (R_{\sigma})_{13}(R_{\sigma})_{12} \label{UR2} \end{equation} if and only if \be \sigma(\alpha,\beta\gamma) = \sigma(\alpha,\beta)\sigma(\alpha,\gamma)\;\;\mbox{\it for all}\;\; \alpha,\beta,\gamma \in \Gamma\,. \label{bich2} \end{equation} 4) Let} $T:\Co\Gas \otimes \Co\Gas \longrightarrow \Co\Gas \otimes \Co\Gas$ {\it be the usual twist mapping and define the function $\sigma^{T} \in {\cal F}(\Gamma \times \Gamma)$ by \be \sigma^{T}(\alpha,\beta) = \sigma(\beta,\alpha)\;\;\mbox{\it for all}\;\; \alpha,\beta \in \Gamma\,. \end{equation} Then we have \be T(R_{\sigma}) = R_{\sigma^{T}}\,. \end{equation}} Proof \\ By direct calculations, using Eqs. \reff{orth} and \reff{compl}. (The notation $(R_{\sigma})_{12}$ etc. is standard, see Section 3.) \\ Finally, we note that \be R_{\sigma}\Delta(h) = (T\Delta(h))R_{\sigma} \label{UR3} \end{equation} for all $h \in \Co\Gas$ and all $\sigma \in {\cal F}(\Gamma \times \Gamma)$. This is obvious since $\Co\Gas$ is commutative and cocommutative. The foregoing results immediately imply the assertion made in the introduction. \\ \noindent {\bf Corollary 2.1} \\ {\it Let $\sigma$ be an arbitrary complex--valued function on $\Gamma \times \Gamma$. Then $R_{\sigma}$ is a universal $R$--matrix for} $\Co\Gas$ {\it if and only if $\sigma$ is a bicharacter on $\Gamma$. Any universal $R$--matrix for} $\Co\Gas$ {\it can uniquely be written in this form, and $R_{\sigma}$ is triangular if and only if $\sigma$ is skew--symmetric in the sense that \be \sigma(\alpha,\beta)\sigma(\beta,\alpha) = 1\;\;\mbox{for all} \;\;\alpha,\beta \in \Gamma\,. \label{skew} \end{equation} The braiding isomorphisms corresponding to a bicharacter $\sigma$ and to the associated universal $R$--matrix $R_{\sigma}$ coincide.} Proof \\ By definition, $R_{\sigma}$ is a universal $R$--matrix if it has an inverse and satisfies the Eqs. \reff{UR1}, \reff{UR2}, and \reff{UR3}. This is the case if and only if $\sigma$ takes its values in $\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}_{\ast}$ and satisfies the Eqs. \reff{bich1} and \reff{bich2}, i.e., if and only if $\sigma$ is a bicharacter on $\Gamma$. Moreover, $R_{\sigma}$ is triangular if and only if \be T(R_{\sigma}) = R^{-1}_{\sigma}\,, \end{equation} i.e., if and only if \be \sigma^{T} = \sigma^{-1}\,. \end{equation} This is exactly Eq. \reff{skew}. The last statement of the corollary has been used to find the correct Ansatz for $R_{\sigma}$\,. \\ {}From a calculational point of view, the parametrization \reff{ans} of the universal $R$--matrices of $\Co\Gas$ is not quite convenient, for it involves the detour over $\Gamma$ and hence a fourfold sum. This drawback can easily be avoided, as follows. \\ \noindent {\bf Proposition 2.2} \\ {\it Let $\Delta_1$ and $\Delta_2$ be two subgroups of $\Gamma^{\ast}$ such that there exists a non--degenerate pairing} \be \tau : \Delta_1 \times \Delta_2 \longrightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}_{\ast}\,. \end{equation} {\it (Since $\Gamma^{\ast}$ is finite this is the case if and only if $\Delta_1$ and $\Delta_2$ are isomorphic.) Let $m$ be the common order of $\Delta_1$ and $\Delta_2$\,. Then the equation \be \sigma(\alpha,\beta) = \frac{1}{m}\sum_{\alpha_i \in \Delta_i}\tau(\alpha_1,\alpha_2) \langle\alpha_1,\alpha\rangle\langle\alpha_2,\beta\rangle \end{equation} defines a bicharacter $\sigma$ on $\Gamma$ and \be R^{\tau} = \frac{1}{m}\sum_{\alpha_i \in \Delta_i}\tau(\alpha_1,\alpha_2)\, \alpha_1 \otimes \alpha_2 \label{Rtau} \end{equation} is the corresponding universal $R$--matrix $R_{\sigma}$ for} $\Co\Gas$. {\it Every bicharacter on $\Gamma$ and hence every universal $R$--matrix for} $\Co\Gas$ {\it can uniquely be obtained in this way.} Proof \\ Needless to say, a non--degenerate pairing $\tau$ of two Abelian groups $\Delta_1$ and $\Delta_2$ is a bicharacter \be \tau : \Delta_1 \times \Delta_2 \longrightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}_{\ast} \end{equation} such that, for any $\alpha_1 \in \Delta_1$\,, the condition \be \tau(\alpha_1,\alpha_2) = 1\;\;\mbox{for all}\;\;\alpha_2 \in \Delta_2 \end{equation} implies that $\alpha_1$ is the unit element of $\Delta_1$\,, and likewise with the roles of $\Delta_1$ and $\Delta_2$ interchanged. The proof of the proposition is quite straightforward, it follows from the standard duality theory for finite Abelian groups. Thus we shall only give a few hints. Let $\sigma$ be a bicharacter on $\Gamma$. Obviously, our task is to investigate the function on $\Gamma^{\ast} \times \Gamma^{\ast}$ defined by \be (\alpha',\beta') \longrightarrow \sum_{\alpha,\beta \in \Gamma}\sigma(\alpha,\beta) \langle\alpha',\alpha\rangle^{\ast}\langle\beta',\beta\rangle^{\ast}\,. \end{equation} Define the (left and right) kernels of $\sigma$ by \begin{eqnarray} N_1 \!\! & = & \!\! \{\alpha \in \Gamma\,|\, \sigma(\alpha,\beta) = 1\;\;\mbox{for all}\;\; \beta \in \Gamma \} \\ N_2 \!\! & = & \!\! \{\beta \in \Gamma\,|\, \sigma(\alpha,\beta) = 1\;\;\mbox{for all}\;\; \alpha \in \Gamma \}\,. \end{eqnarray} Then $\sigma$ induces a non--degenerate pairing $\tilde{\sigma}$ of $\Gamma/N_1$ and $\Gamma/N_2$\,, hence the orders of these groups coincide. We denote them by $m$. Define the subgroups $\Delta_1$ and $\Delta_2$ of $\Gamma^{\ast}$ by \be \Delta_i = \{\gamma' \in \Gamma^{\ast}\,|\, \langle\gamma',\gamma_i\rangle = 1\;\;\mbox{for all}\;\; \gamma_i \in N_i \}\,. \end{equation} It is well--known that $\Delta_i$ can be identified with the character group of $\Gamma/N_i$\,, thus $\tilde{\sigma}$ induces a non--degenerate pairing $\tau$ of $\Delta_1$ and $\Delta_2$\,. This is the one we are looking for. To be more explicit, define the complex--valued function $\sigma'$ on $\Gamma^{\ast} \times \Gamma^{\ast}$ by \be \sigma'(\alpha',\beta') = \frac{m}{n^2}\sum_{\alpha,\beta \in \Gamma}\sigma(\alpha,\beta) \langle\alpha',\alpha\rangle^{\ast}\langle\beta',\beta\rangle^{\ast}\,, \end{equation} where the normalization factor \be \frac{m}{n^2} = \frac{1}{m}\frac{1}{(n/m)^2} \end{equation} has been chosen conveniently. Using the duality theory for finite Abelian groups it can be shown that $\sigma'(\alpha',\beta')$ vanishes whenever $\alpha' \notin \Delta_1$ or $\beta' \notin \Delta_2$\,. Let $\tau$ be the restriction of $\sigma'$ onto $\Delta_1 \times \Delta_2$\,. Then $\tau$ is the pairing described abstractly above and has all the properties stated in the proposition. \\ \noindent {\bf Remark} \\ The universal $R$--matrices given by Eq. \reff{Rtau} with $\Delta_1 = \Delta_2 = \Gamma^{\ast}$ and $\tau$ a non--degenerate commutation factor on $\Gamma^{\ast}$ have also been obtained in Ref. [9]. \\ \noindent {\bf Example} \\ Let $n \ge 1$ be an integer, let $\Gamma_n$ be the (multiplicative) group of all $n$th roots of unity, and let $Z_n$ be the ring of integers modulo $n$\,. Obviously, for any $\alpha \in \Gamma_n$ and $r \in Z_n$\,, the power $\alpha^r$ has a well--defined meaning. Define \be \omega = e^{2\pi i/n} \end{equation} and let $\gamma$ be a primitive $n$th root of unity, i.e., a generator of the group $\Gamma_n$ (for example, $\gamma = \omega$). Let $\chi : \Gamma_n \to \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}_{\ast}$ be the character defined by \be \langle\chi,\gamma^r\rangle = \omega^r\;\;\mbox{for all}\;\;r \in Z_n\,. \end{equation} Then there exists a unique isomorphism of $\Gamma_n$ onto $\Gamma_{n}^{\ast}$ which maps $\gamma$ onto $\chi$\,, and we could use this isomorphism to identify $\Gamma_{n}^{\ast}$ with $\Gamma_n$\,. The canonical pairing is given by \be \langle\chi^r,\gamma^s\rangle = \omega^{rs}\;\;\mbox{for all}\;\;r,s \in Z_n\,, \end{equation} and an arbitrary bicharacter on $\Gamma_n$ has the form \be \sigma_k(\gamma^r,\gamma^s) = \omega^{krs}\;\;\mbox{for all}\;\;r,s \in Z_n\,, \end{equation} with some element $k \in Z_n$\,. According to Eq. \reff{ans}, the universal $R$--matrix $R_k \in \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}\Gamma_{n}^{\ast} \otimes \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}\Gamma_{n}^{\ast}$ corresponding to $\sigma_k$ is given by \be R_k = \frac{1}{n^2}\sum_{p,q,r,s \in Z_n}\omega^{kpq}\omega^{-pr}\omega^{-qs} \chi^r \otimes \chi^s\,. \end{equation} Obviously, part of this sum can be carried out. Consider, for a moment, $k$ as an element of $\{1,2,\ldots,n\}$ and let $d$ be the greatest common divisor of $k$ and $n$\,. Let $\ell$ be the inverse of $k/d$ in the ring $Z_{n/d}$\,. Then \be R_k = \frac{d}{n}\sum_{x,\,y \in Z_{n/d}}(\omega^d)^{-\ell xy} (\chi^d)^x \otimes (\chi^d)^y\,. \label{Rex} \end{equation} Note that in this case the subgroups $\Delta_1$ and $\Delta_2$ of $\Gamma_{n}^{\ast}$ introduced in Proposition 2.2 are equal to \be \Delta_1 = \Delta_2 = \{(\chi^d)^x\,|\, x \in Z_{n/d}\} \end{equation} and are isomorphic to $\Gamma_{n/d}$\,. The universal $R$--matrix \reff{Rex} with $k=1$ (and hence $d=1$, $\ell=1$) has already been used in Ref. [10], the special case $n=2$ (related to supersymmetry) is contained in Ref. [11]. \\ We would now like to interpret our results from a dual point of view. However, this necessitates some preparation. Consequently, we shall first comment on the duality of Hopf algebras and return to our main topic not until Section 4. \section{Duality of Hopf algebras and \newline coquasitriangular structures} \setcounter{equation}{0} The results of the present section are well--known to experts in Hopf algebra theory. They are included for the convenience of the reader and for later reference. Let $A$ and $H$ be two bialgebras [2] and let \be \varphi : A \times H \longrightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}} \end{equation} be a bilinear form on $A \times H$. Then there exists a unique bilinear form \be \varphi^{\ot2}: (A \otimes A) \times (H \otimes H) \longrightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}} \end{equation} such that \be \varphi^{\ot2}(a \otimes a',h \otimes h') = \varphi(a,h)\,\varphi(a',h') \label{quad} \end{equation} for all $a,a' \in A$ and $h,h' \in H$. The form $\varphi$ is called a Hopf (or bialgebra) pairing of $A$ and $H$ if the following relations are satisfied: \begin{eqnarray} \varphi^{\ot2}(\Delta_{A}(a),h \otimes h') \!\! & = & \!\! \varphi(a,hh') \\ \varphi^{\ot2}(a \otimes a',\Delta_{H}(h)) \!\! & = & \!\! \varphi(aa',h) \end{eqnarray} \vspace{-6ex} \begin{eqnarray} \varphi(a,1_{H}) \!\! & = & \!\! \varepsilon_{A}(a) \\ \varphi(1_{A},h) \!\! & = & \!\! \varepsilon_{H}(h) \end{eqnarray} for all $a,a' \in A$ and all $h,h' \in H$. Here, $\Delta_A$, $\varepsilon_A$, and $1_A$ denote the coproduct, the counit, and the unit element of $A$\,, and $\Delta_H$, $\varepsilon_H$, and $1_H$ have the analogous meaning. {}From the point of view of dual Hopf algebras (see Ref. [2] and below) the concept of a Hopf pairing is very natural and has been used by many authors. The first reference I know of is Ref. [12]. The prototype of Hopf pairings can be constructed as follows. Let $H$ be a bialgebra, let $H^{\ast}$ be the vector space dual to $H$, and let $H^{\circ} \subset H^{\ast}$ be the set of all linear forms on $H$ which vanish on an ideal of $H$ of finite codimension. It is easy to see that $H^{\circ}$ is a subspace of $H^{\ast}$. Ref. [2] contains various characterizations of the elements of $H^{\circ}$. In particular, let $m:H \otimes H \rightarrow H$ be the product mapping of $H$ and let $m^{t}: H^{\ast} \rightarrow (H \otimes H)^{\ast}$ be its transpose. Recall that $H^{\ast} \otimes H^{\ast}$ is canonically embedded in $(H \otimes H)^{\ast}$. Then an element $f \in H^{\ast}$ belongs to $H^{\circ}$ if and only if $m^t(f) \in H^{\ast} \otimes H^{\ast}$. Now let $\varphi$ be the restriction of the canonical pairing \be \langle\,\;,\;\rangle:H^{\ast} \times H \longrightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}} \end{equation} onto $H^{\circ} \times H$. Then there exists a unique bialgebra structure on $H^{\circ}$ such that $\varphi$ is a Hopf pairing of $H^{\circ}$ and $H$ (for a proof, see Ref. [2]). We call the bialgebra $H^{\circ}$ the Hopf dual of $H$. Using this result, the Hopf pairings can be characterized as follows. Let us use the notation introduced at the beginning of this section. It is well--known that, for any bilinear form $\varphi$ on $A \times H$, there exist two linear mappings \be \varphi_{\ell}:A \longrightarrow H^{\ast} \end{equation} \be \varphi_{r}:H \longrightarrow A^{\ast} \end{equation} such that \be \langle\varphi_{\ell}(a),h\rangle = \langle\varphi_{r}(h),a\rangle = \varphi(a,h) \end{equation} for all $a \in A$ and $h \in H$ (obviously, they are uniquely determined by this equation). Then we can prove the following proposition. \\ \noindent {\bf Proposition 3.1} \\ {\it We use the notation introduced above. Then the following statements are equivalent: \\ 1) The bilinear form $\varphi$ is a Hopf pairing of $A$ and $H$. \\ 2) We have $\varphi_{\ell}(A) \subset H^{\circ}$, and the map of $A$ into $H^{\circ}$ induced by $\varphi_{\ell}$ is a bialgebra homomorphism. \\ 3) We have $\varphi_{r}(H) \subset A^{\circ}$, and the map of $H$ into $A^{\circ}$ induced by $\varphi_{r}$ is a bialgebra homomorphism.} Proof \\ The only non--trivial part of this proposition is that, for any Hopf pairing, we have \be \varphi_{\ell}(A) \subset H^{\circ}\;\;\; \mbox{and}\;\;\;\varphi_{r}(H) \subset A^{\circ}\,, \end{equation} and this follows from the characterization of $H^{\circ}$ and $A^{\circ}$ recalled above. The rest can be checked by straightforward calculations. \\ Suppose now that $H$ is even a Hopf algebra, and let $S_H$ be the antipode of $H$. Then it is easy to see that the transpose $S_{H}^{t}$ of $S_H$ maps $H^{\circ}$ into itself and that the map of $H^{\circ}$ into itself induced by $S_{H}^{t}$ is an antipode for $H^{\circ}$. Thus if $H$ is a Hopf algebra, then $H^{\circ}$ is likewise. We now can prove the following corollary to Proposition 3.1. \\ \noindent {\bf Corollary 3.1} \\ {\it We use the notation introduced at the beginning of this section, but suppose in addition that $A$ and $H$ are Hopf algebras, with antipodes $S_A$ and $S_H$, respectively. If $\varphi$ is a Hopf pairing of $A$ and $H$, then \be \varphi(S_{A}(a),h) = \varphi(a,S_{H}(h)) \end{equation} for all $a \in A$ and $h \in H$.} Proof \\ It is well--known [2] that a bialgebra homomorphism of a Hopf algebra into another one is, in fact, a Hopf algebra homomorphism, i.e., it intertwines the antipodes. Thus the corollary follows from Proposition 3.1. \\ Our next task is to dualize the notion of a universal $R$--matrix [1]. Let $H$ be a bialgebra. An element $R \in H \otimes H$ is called a universal $R$--matrix for $H$ if it has the following four properties: \be R\;\;\mbox{is invertible in}\;\;H \otimes H \label{URM1} \end{equation} \vspace{-5.5ex} \be (T\Delta_{H}(h))R = R\,\Delta_{H}(h)\;\;\mbox{for all}\;\;h \in H \label{URM2}\end{equation} \vspace{-7ex} \begin{eqnarray} (\Delta_{H} \otimes id_{H})(R) \!\! & = & \!\! R_{13}R_{23} \label{URM3} \\ (id_{H} \otimes \Delta_{H})(R) \!\! & = & \!\! R_{13}R_{12}\,. \label{URM4} \end{eqnarray} Our notation is standard: \be T:H \otimes H \longrightarrow H \otimes H \end{equation} is the linear twist mapping given by \be T(h \otimes h') = h' \otimes h\;\;\mbox{for all}\;\;h,h' \in H \end{equation} and the elements \be R_{12}\,,\:R_{23}\,,\:R_{13} \in H \otimes H \otimes H \end{equation} are defined by \be R_{12} = R \otimes 1_H \end{equation} \be R_{23} = 1_H \otimes R \end{equation} \be R_{13} = (T \otimes id_H)(R_{23}) = (id_H \otimes T)(R_{12})\,, \end{equation} where $1_H$ denotes the unit element of $H$. In order to dualize these four properties, we assume that $H$ is the Hopf dual of a bialgebra $A$\,, \be H = A^{\circ}\,. \end{equation} Since $A$ and hence $A \otimes A$ are coalgebras, their duals $A^{\ast}$ and $(A \otimes A)^{\ast}$ are algebras, and $A^{\circ}$ is a subalgebra of $A^{\ast}$. Actually, we have the following chain of canonical injective algebra homomorphisms: \be A^{\circ} \otimes A^{\circ} \longrightarrow A^{\ast} \otimes A^{\ast} \longrightarrow (A \otimes A)^{\ast}\,. \label{inj} \end{equation} Thus, since $R$ is an element of $A^{\circ} \otimes A^{\circ}$, it can also be considered as an element of $(A \otimes A)^{\ast}$. On the other hand, due to the familiar properties of tensor products of vector spaces, $(A \otimes A)^{\ast}$ can be canonically identified with the space $B(A,A)$ of all bilinear forms on $A \times A$\,. In particular, $B(A,A)$ inherits from $(A \otimes A)^{\ast}$ the structure of an associative algebra with a unit element. The multiplication is the so--called convolution. It is denoted by an asterisk $\ast$ and is defined as follows. Let $\psi,\,\psi'$ be two elements of $B(A,A)$ and let $a,b \in A$\,. If $\Delta_A$ is the coproduct of $A$ and if \be \Delta_{A}(a) = \sum_{i}a^1_i \otimes a^2_i \label{dela} \end{equation} \be \Delta_{A}(b) = \sum_{j}b^1_j \otimes b^2_j \label{delb} \end{equation} with $a^1_i,a^2_i,b^1_j,b^2_j \in A$\,, we have \be(\psi \ast \psi')(a,b) = \sum_{i,j}\psi(a^1_i,b^1_j) \,\psi'(a^2_i,b^2_j) \,. \label{conv} \end{equation} The unit element is the bilinear form $\varepsilon_B$ on $A \times A$ defined by \be \varepsilon_B(a,b) = \varepsilon_A(a)\varepsilon_A(b) = \varepsilon_A(ab) \end{equation} for all $a,b \in A$\,, where $\varepsilon_A$ denotes the counit of $A$\,. Now if \be R = \sum_{i}R^{1}_{i} \otimes R^{2}_{i} \end{equation} is an arbitrary element of $A^{\circ} \otimes A^{\circ}$, the corresponding bilinear form $\varrho \in B(A,A)$ is given by \be \varrho(a,b) = \sum_{i}\langle R^{1}_{i},a\rangle\langle R^{2}_{i},b\rangle \end{equation} for all $a,b \in A$\,, where $\langle\,\;,\;\rangle$ denotes the canonical pairing of $A^{\ast}$ and $A$\,. If $R$ is invertible in $A^{\circ} \otimes A^{\circ}$, then all the more in $B(A,A)$. This means that there exists a bilinear form $\varrho' \in B(A,A)$ such that \be \varrho \ast \varrho' = \varrho' \ast \varrho = \varepsilon_B \,. \label{inv} \end{equation} Note that if $A$ is finite--dimensional, then the homomorphisms in Eq. \reff{inj} are bijective, and hence $R$ is invertible in $A^{\circ} \otimes A^{\circ}$ if and only if a bilinear form $\varrho'$ of the type described above exists. In general, however, we cannot expect that the invertibility of $\varrho$ in $B(A,A)$ implies the invertibility of $R$ in $A^{\circ} \otimes A^{\circ}$. The remaining conditions \reff{URM2} -- \reff{URM4} can also be reformulated in terms of certain convolution products involving $\varrho$\,. Because of lack of space, we prefer to present the results more explicitly, as follows. Condition \reff{URM2} is satisfied if (and, if $A^{\circ}$ separates the elements of $A$\,, only if) the following equation holds for all $a,b \in A$\,: \be \sum_{i,j}b^{1}_{j} a^{1}_{i} \varrho(a^{2}_{i},b^{2}_{j}) = \sum_{i,j}\varrho(a^{1}_{i},b^{1}_{j}) a^{2}_{i} b^{2}_{j} \label{comm} \end{equation} (here and in the subsequent equations we are using the notation introduced in Eqs. \reff{dela}, \reff{delb}). Moreover, Eq. \reff{URM3} is satisfied if and only if \be \varrho(bc,a) = \sum_{i}\varrho(b,a^1_i)\varrho(c,a^2_i) \label{rho3} \end{equation} for all $a,b,c \in A$\,, and Eq. \reff{URM4} holds if and only if \be \varrho(a,bc) = \sum_{i}\varrho(a^1_i,c)\varrho(a^2_i,b) \label{rho4} \end{equation} for all $a,b,c \in A$\,. The foregoing results motivate the following definition. \\ \noindent {\bf Definition 3.1} \\ {\it Let $A$ be a bialgebra. A bilinear form $\varrho$ on $A \times A$ is said to define a coquasitriangular (or dual quasitriangular, or braiding) structure on $A$ if it is convolution invertible (i.e., if it has an inverse in the sense of Eq. \reff{inv}) and if the relations \reff{comm}, \reff{rho3}, and \reff{rho4} are satisfied. A bialgebra endowed with a coquasitriangular structure will be called coquasitriangular.} \\ Apparently, dual quasitriangular structures have first been introduced in Ref. [13] and shortly afterwards by various authors. In particular, I refer the reader to Ref. [14] (note that the coquasitriangular structures discussed here are the right braiding structures defined there). Visibly, Eq. \reff{comm} is some sort of generalized commutation relation. In fact, the famous RTT relations of Ref. [15] are of this type, and the corresponding bi\-algebras $A(R)$ (with $R$ an invertible solution of the Yang--Baxter equation) are coquasitriangular. The Eqs. \reff{rho3} and \reff{rho4} can be rewritten as follows. Let $\varrho^{\otimes 2}$ be the bilinear form on $A \otimes A$ defined by \be \varrho^{\otimes 2}(a \otimes a',b \otimes b') = \varrho(a,b)\varrho(a',b') \end{equation} for all $a,a',b,b' \in A$ (see Eq. \reff{quad}). Then Eq. \reff{rho3} is equivalent to \be \varrho^{\otimes 2}(b \otimes c, \Delta_{A}(a)) = \varrho(bc,a)\,, \label{rho3mod} \end{equation} and Eq. \reff{rho4} is equivalent to \be \varrho^{\otimes 2}(T\Delta_{A}(a), b \otimes c) = \varrho(a, bc)\,, \label {rho4mod} \end{equation} (where, once again, $T$ denotes the twist mapping) or, what amounts to the same, to \be \varrho^{\otimes 2}(\Delta_{A}(a), c \otimes b) = \varrho(a,bc)\,. \label{rho4modd} \end{equation} As is well--known, there is a real host of properties and relations satisfied by a universal $R$--matrix. Thus one should expect the same for coquasitriangular structures. Of course, due to the well--known problems with duality for infinite--dimensional vector spaces, we cannot simply prove these latter results by duality. Rather, we use duality to find out which properties should be true, and then try to prove these directly. Let us give some examples. In the following, $R$ denotes a universal $R$--matrix for a bialgebra $H$ and $\varrho$ denotes a coquasitriangular structure for a bialgebra $A$\,. It is obvious that with $R$ also $T(R^{-1}) = T(R)^{-1}$ is a universal $R$--matrix for $H$ (as before, $T$ denotes the appropriate twist mapping, here the one of $H \otimes H$ onto itself). Dually, with $\varrho$ also $\varrho' \!\stackrel{\textstyle{\hst{0ex} \atop \circ}}{\hst{0ex}}\! T = (\varrho \!\stackrel{\textstyle{\hst{0ex} \atop \circ}}{\hst{0ex}}\! T)'$ is a coquasitriangular structure for $A$\,, where the prime denotes the convolution inverse. A coquasitriangular structure $\varrho$ is said to be cotriangular if $\varrho' \!\stackrel{\textstyle{\hst{0ex} \atop \circ}}{\hst{0ex}}\! T = \varrho$\,. It is known that \be (\varepsilon_H \otimes id_H)(R) = 1_H \end{equation} \be (id_H \otimes \varepsilon_H)(R) = 1_H \end{equation} (in these equations, $\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}} \otimes H$ and $H \otimes \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}$ have been canonically identified with $H$). Dually, it can be shown that \be \varrho(1_A,a) = \varrho(a,1_A) = \varepsilon_A(a) \label{rhounit} \end{equation} for all $a \in A$ (for example, apply Eq. \reff{rho4} with $c = 1_A$ and use that $\varrho$ has a convolution inverse). Finally, let us assume that $H$ and $A$ are Hopf algebras. Then it is known that \begin{eqnarray} (S_H \otimes id_H)(R) \!\! & = & \!\! R^{-1} \\ (id_H \otimes S_H)(R^{-1}) \!\! & = & \!\! R\,. \end{eqnarray} Dually, it can be shown that \begin{eqnarray} \varrho(S_A(a),b) \!\! & = & \!\! \varrho'(a,b) \label{rhoS1} \\ \varrho'(a,S_A(b)) \!\! & = & \!\! \varrho(a,b) \end{eqnarray} for all $a,b \in A$\,, where, as before, $\varrho'$ is the convolution inverse of $\varrho$\,. In particular, we have \be \varrho(S_A(a),S_A(b)) = \varrho(a,b) \end{equation} for all $a,b \in A$\,. Conversely, if $\varrho$ is a bilinear form on a Hopf algebra $A$ satisfying the Eqs. \reff{comm} -- \reff{rho4} and \reff{rhounit}, then the bilinear form $\varrho'$ defined by Eq. \reff{rhoS1} is a convolution inverse of $\varrho$ and hence $\varrho$ defines a coquasitriangular structure on $A$\,. Let us next make contact with our discussion of Hopf pairings. Let $A$ be a bi\-algebra and let $A^{cop}$ (resp. $A^{aop}$) be the bialgebra which, considered as an algebra (resp. a coalgebra) coincides with $A$\,, but which, considered as a coalgebra (resp. an algebra) has the structure opposite to that of $A$\,. Then the Eqs. \reff{rho3mod} -- \reff{rho4modd} and \reff{rhounit} show that a coquasitriangular structure $\varrho$ of $A$ can be defined to be a Hopf pairing $A^{cop} \times A \rightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}$ (or, equivalently, a Hopf pairing $A \times A^{aop} \rightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}$) which has a convolution inverse in the sense of Eqs. \reff{conv}, \reff{inv} and satisfies the generalized commutation relations \reff{comm}. In particular, it follows from Proposition 3.1 that, for any $b \in A$\,, the linear forms $a \rightarrow \varrho(a,b)$ and $a \rightarrow \varrho(b,a)$ on $A$ belong to $A^{\circ}$. This result is highly welcome. In our dualization process, we started from a universal $R$--matrix $R \in A^{\circ} \otimes A^{\circ}$ and embedded $A^{\circ} \otimes A^{\circ}$ into $(A \otimes A)^{\ast} \simeq B(A,A)$. For infinite--dimensional bialgebras, $(A \otimes A)^{\ast}$ is terribly much larger than $A^{\circ} \otimes A^{\circ}$. On the other hand, a certain enlargement of $A^{\circ} \otimes A^{\circ}$ is reasonable, since for the most interesting examples of infinite--dimensional quasitriangular Hopf algebras $H$, the universal $R$--matrix does not belong to $H \otimes H$ but to a certain completion of it. Our last result shows that $\varrho$ seems to be not as far from $A^{\circ} \otimes A^{\circ}$ as one might have feared. Thus one could hope to find a subbialgebra $U$ of $A^{\circ}$ associated to $\varrho$ and a suitable completion of $U \otimes U$ in $B(A,A)$ containing $\varrho$\,. Such a programme has been carried out in Ref. [14]. Finally, let us recall that, for any coalgebra $C$\,, the right $C$--comodules may also be regarded as left $C^{\ast}$--modules (where $C^{\ast}$ denotes the associative algebra dual to $C$)[2]. In view of the discussion of the present section we expect that a coquasitriangular structure on a bialgebra $A$ can be used to convert the class of right $A$--comodules into a braided monoidal category. This can in fact be done (actually, this is the starting point of Ref. [14]). We do not want to go into detail here but only show how the corresponding braiding isomorphisms are defined. Thus let $V$ and $W$ be two right $A$--comodules, with structure maps \begin{eqnarray} \delta_{V} \!\!\!& : &\!\!\! V \longrightarrow V \otimes A \\ \delta_{W} \!\!\!& : &\!\!\! W \longrightarrow W \otimes A \,. \end{eqnarray} To define the braiding isomorphism \be \psi_{V,W} : V \otimes W \longrightarrow W \otimes V \:, \end{equation} choose $x \in V$, $y \in W$, and write \begin{eqnarray} \delta_{V}(x) \!\! & = & \!\! \sum_{i}x_i \otimes a_i \\ \delta_{W}(y) \!\! & = & \!\! \sum_{j}y_j \otimes b_j \end{eqnarray} with $x_i \in V$, $y_j \in W$\,; $a_{i},b_{j} \in A$\,. Then we set \be \psi_{V,W}(x \otimes y) = \sum_{i,j}\varrho(a_{i},b_{j})\, y_j \otimes x_i \;. \label{cobraid} \end{equation} We close this section by two remarks. For reasons of uniformity, we have formulated the results of this section over the base field $\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}$\,. Of course, everything holds for arbitrary base fields of any characteristic. Moreover, by systematically using the techniques supplied by Ref. [3], these results can easily be generalized to the graded case, with arbitrary Abelian groups of degrees and arbitrary commutation factors. \section{Coquasitriangular structures for the group \newline Hopf algebra of an Abelian group} \setcounter{equation}{0} We now are ready to rederive the results of Section 2 from the dual point of view. Moreover, we are going to see that the dual picture is applicable also to infinite Abelian groups. To begin with we still assume that the group $\Gamma$ is finite. Then the group Hopf algebras $\Co\Ga$ and $\Co\Gas$ are (Hopf) dual to each other. In fact, since $\Gamma$ is a basis of $\Co\Ga$ and $\Gamma^{\ast}$ a basis of $\Co\Gas$, there is a unique bilinear form $\Co\Gas \times \Co\Ga \rightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}$ whose restriction to $\Gamma^{\ast} \times \Gamma$ is equal to the group pairing $\Gamma^{\ast} \times \Gamma \rightarrow \mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}}$\,, and it is easy to check that this bilinear form is a non--degenerate Hopf pairing. Thus there is no problem in denoting these two pairings by the same symbol $\langle\,\;,\;\rangle$. The result above is well--known, it agrees with the fact that the dual of the group Hopf algebra $\mbox{C\hst{-0.44em}\vl{0.1ex}{1.4ex}\hst{0.44em}} G$ of a finite group $G$ is canonically isomorphic to the Hopf algebra ${\cal F}(G)$ of complex--valued functions on $G$\,. In fact, the characters of $\Gamma$ form a basis of ${\cal F}(\Gamma)$. Of course, all this is closely related to Fourier transformation on $\Gamma$ and $\Gamma^{\ast}$. According to Section 3 the foregoing result implies that the universal $R$--matrices for $\Co\Gas$ are in bijective correspondence with the coquasitriangular structures on $\Co\Ga$. More precisely, consider the chain \reff{inj} of injective algebra homomorphisms, with $A = \Co\Ga$. In the present case these are, in fact, algebra isomorphisms, and according to the foregoing result we can (and will) identify $(\Co\Ga)^{\ast}$ with $\Co\Gas$. On the other hand $(\Co\Ga \otimes \Co\Ga)^{\ast}$ is canonically isomorphic to $B(\Co\Ga,\Co\Ga)$, the algebra of bilinear forms on $\Co\Ga \times \Co\Ga$. Since $\Gamma$ is a basis of $\Co\Ga$, the bilinear forms on $\Co\Ga \times \Co\Ga$ are uniquely determined by their values on $\Gamma \times \Gamma$, which shows that the vector space $B(\Co\Ga,\Co\Ga)$ can be identified with ${\cal F}(\Gamma \times \Gamma)$, the vector space of complex--valued functions on $\Gamma \times \Gamma$. Summarizing, we have the following chain of vector space isomorphisms: \be \Co\Gas \otimes \Co\Gas \longrightarrow (\Co\Ga \otimes \Co\Ga)^{\ast} \longrightarrow B(\Co\Ga,\Co\Ga) \longrightarrow {\cal F}(\Gamma \times \Gamma)\,. \end{equation} We know that the first of these maps is an algebra homomorphism and (by definition of the convolution) so is the second. A look at Eq. \reff{conv} shows that the third map is an algebra homomorphism as well. Furthermore, it is easy to check that the inverse mapping \be {\cal F}(\Gamma \times \Gamma) \longrightarrow \Co\Gas \otimes \Co\Gas \end{equation} is just the mapping $\sigma \rightarrow R_{\sigma}$ mentioned in part 1) of Proposition 2.1. Thus all we have to do is to determine those bilinear forms $\varrho$ on $\Co\Ga \times \Co\Ga$ which define a coquasitriangular structure on $\Co\Ga$. Since $\Gamma$ is Abelian, the condition \reff{comm} is trivially satisfied. Obviously, the remaining conditions (convolution invertibility and Eqs. \reff{rho3}, \reff{rho4}) are fulfilled if and only if the restriction of $\varrho$ to $\Gamma \times \Gamma$ (i.e., the element of ${\cal F}(\Gamma \times \Gamma)$ corresponding to $\varrho$) is a bicharacter on $\Gamma$. Thus we have rederived part 1) of Proposition 2.1 and the non--trivial parts of Corollary 2.1. {\it Let us now drop the assumption that $\Gamma$ is finite.} Then the coquasitriangular structures on $\Co\Ga$ can be characterized as before: A bilinear form $\varrho$ on $\Co\Ga \times \Co\Ga$ defines a coquasitriangular structure on $\Co\Ga$ if and only if its restriction $\sigma$ to $\Gamma \times \Gamma$ is a bicharacter of $\Gamma$, and $\varrho$ is cotriangular if and only if $\sigma$ is a commutation factor on $\Gamma$. Of course, we should now answer the question of how graded multilinear algebra (which has been the point of departure for our round trip) can be formulated in the present setting. The answer is suggested by the remarks at the end of Section 3: The $\Co\Ga$--comodules should take the role of the $\Co\Gas$--modules. In fact, it is well--known that $\Gamma$--graded vector spaces and $\Co\Ga$--comodules are essentially the same thing. Indeed, let $V$ be a right $\Co\Ga$--comodule and let \be \delta : V \longrightarrow V \otimes \Co\Ga \end{equation} be its structure mapping. If we define, for all $\gamma \in \Gamma$, \be V_{\gamma} = \{x \in V \,|\, \delta(x) = x \otimes \gamma \}\,, \end{equation} then, obviously, $V_{\gamma}$ is a subspace of $V$, and it is easy to see that these subspaces form a $\Gamma$--gradation of $V$\,: \be V = \bigoplus_{\gamma \in \Gamma} V_{\gamma}\,. \end{equation} Conversely, if $(V_{\gamma})_{\gamma \in \Gamma}$ is a $\Gamma$--gradation of a vector space $V$, there exists a unique linear mapping \be \delta : V \longrightarrow V \otimes \Co\Ga \end{equation} such that \be \delta(x) = x \otimes \xi \end{equation} for all $x \in V_{\xi}$\,, $\xi \in \Gamma$, and this mapping defines on $V$ the structure of a right $\Co\Ga$--comodule. Obviously, these transitions from right $\Co\Ga$--comodules to $\Gamma$--graded vector spaces and vice versa are inverse to each other. Moreover, the homomorphisms of $\Co\Ga$--comodules are exactly the homomorphisms of $\Gamma$--graded vector spaces. Now let $\varrho$ define a coquasitriangular structure on $\Co\Ga$ and let $\sigma$ be the restriction of $\varrho$ to $\Gamma \times \Gamma$. Consider two $\Gamma$--graded vector spaces, i.e., right $\Co\Ga$--comodules, denoted by $V$ and $W$, say. Then the braiding isomorphism \be \psi_{V,W} : V \otimes W \longrightarrow W \otimes V \end{equation} corresponding to $\varrho$ (see Eq. \reff{cobraid}) is given by \be \psi_{V,W}(x \otimes y) = \sigma(\xi,\eta)\,y \otimes x \end{equation} for all $x \in V_{\xi}$\,, $y \in W_{\eta}$\,; $\xi,\eta \in \Gamma$. This is exactly the formula familiar from graded multilinear algebra [3]. Needless to say, there is a lot of other details which should be checked before we can be sure that the graded multilinear algebra as developed in Ref. [3] and the comodule picture described here (with $\varrho$ cotriangular) are really equivalent. However, I do not want to embark on this boring exercise. Instead, I would like to close this section by the remark that we have come across one more example of a well--known fact: In the infinite--dimensional case, coalgebras and comodules are easier to handle than algebras and modules. \section{Discussion} \setcounter{equation}{0} In the present work we have described how the graded multilinear algebra as developed in Ref. [3] can be understood in the language of triangular or cotriangular Hopf algebras. We may now proceed to apply the theory of Hopf algebras to the study of algebras living in the graded multilinear category. In particular, we could apply Majid's theory of bosonization [16] and try to reduce (or, at least, to relate) the theory of Lie colour algebras to the theory of usual Lie algebras. (Roughly speaking, bosonization is a procedure which converts a Hopf algebra living in a category with non--trivial braiding into a usual Hopf algebra.) The idea to try this is not new. It is used in supersymmetry by introducing Grassmann variables, and it is also applied in Ref. [5], where the present author has reduced the theory of Lie colour algebras to the theory of graded Lie (super)algebras. More recent and closer to the present work and to the bosonization procedure are the Refs. [17,\,9]. In Ref. [17] the authors prove, in the context of coquasitriangular Hopf algebras, a general bosonization theorem and use it to derive Schur's double centralizer theorem for cotriangular Hopf algebras (hence, in particular, for Lie colour algebras). In Ref. [9] the author constructs a bosonized version of the enveloping algebra of a Lie colour algebra and determines its universal $R$--matrix. Of course, there is no need to restrict attention to the triangular case, i.e., to commutation factors. In fact, allowing for arbitrary bicharacters, the graded vector spaces are the objects of a really braided monoidal category, i.e., one in which the braid groups really come into play. On the other hand, these categories are still very close to the classical case, hence they might be used to become acquainted with the braided situation. For example, one could try to define generalized Lie algebras living in these categories. This has in fact been done in a recent work by Pareigis [18]. Using the tensor operator techniques developed in Refs. [19,\,20] one might then hope to construct interesting quantum spin chain Hamiltonians which are invariant under these algebras.\\[2ex] \noindent {\bf Acknowledgements}\,\, A major part of the present work has been carried out during two visits of the author to the Erwin Schr\"{o}dinger Institut in Vienna. The kind invitations by Harald Grosse and the hospitality extended to the author in the ESI are gratefully acknowledged. \\ \newpage \noindent {\large\bf References} \begin{enumerate} \item V. Drinfeld, J. Sov. Math. {\bf 41}, 898 (1988) (expanded version of a report to the International Congress of Mathematicians, Berkeley 1986). \item M.E. Sweedler, {\it Hopf algebras}, W.A. Benjamin, New York (1969). \item M. Scheunert, J. Math. Phys. {\bf 24}, 2658 (1983). \item V. Rittenberg and D. Wyler, Nucl. Phys. B {\bf 139}, 189 (1978). \item M. Scheunert, J. Math. Phys. {\bf 20}, 712 (1979). \item S. Mac Lane, {\it Categories for the working mathematician}, Springer, New York, Heidelberg, Berlin (1971). \item A. Joyal and R. Street, Braided monoidal categories, Mathematics Reports 860081, Macquarie University (1986), and Adv. in Math. {\bf 102}, 20 (1993). \item S. Majid, Int. J. Mod. Phys. A {\bf 5}, 1 (1990). \item D.S. McAnally, Lett. Math. Phys. {\bf 33}, 249 (1995). \item S. Majid, in {\it Spinors, twistors, Clifford algebras, and quantum deformations,} Proceedings of the 2\,nd Max Born Symposium, Wroclaw 1992, edited by Z. Oziewicz et al., Kluwer, pp. 327--336. \item D. Radford, J. Algebra {\bf 141}, 354 (1991). \item M. Takeuchi, {\it The \#--product of group sheaf extensions applied to Long's theory of dimodule algebras,} Algebra--Berichte, Mathematisches Institut der Universit\"{a}t M\"{u}nchen, Nr. 34, Verlag Uni--Druck, M\"{u}nchen (1977). \item S. Majid, in {\it Quantum probability and related topics VI,} Trento 1989, proceedings, edited by L. Accardi et al., World Scientific, Singapore (1991), pp. 333--358. \item R.G. Larson and J. Towber, Comm. Algebra {\bf 19}, 3295 (1991). \item N.Yu. Reshetikhin, L.A. Takhtadzhyan, and L.D. Faddeev, Leningrad Math. J. {\bf 1}, 193 (1990). \item S. Majid, J. Algebra, {\bf 163}, 165 (1994). \item D. Fischman and S. Montgomery, J. Algebra {\bf 168}, 594 (1994). \item B. Pareigis, On Lie algebras in braided categories, Mathematisches Institut der Universit\"{a}t M\"{u}nchen, preprint, April 1995. \item M. Scheunert, in {\it Generalized symmetries in physics,} Clausthal 1993, proceedings, edited by H.--D. Doebner, V.K. Dobrev, and A.G. Ushveridze, World Scientific (1994), pp. 77--89. \item M. Scheunert, Int. J. Mod. Phys. B {\bf 8}, 3655 (1994). \end{enumerate} \end{document}
\section{Introduction} Few-body problems of interacting particles have vital importance in all branches of physics from hadron to celestial levels. The main interest in the few-body problems lies in, e.g., finding an accurate solution for the system, testing the equation of motion and the conservation laws and symmetries, or looking for unknown interactions governing the system. \par\indent The investigation of few-nucleon systems interacting via realistic forces has always been in the center of the interest. Considerable effort has been exerted to obtain accurate ground-state properties of the few-nucleon systems with Faddeev-Yakubovsky (FY) \cite{payne,faddeev,glockb}, variational \cite{atms,hypers,crcbgv}, variational Monte Carlo (VMC) \cite{mcv,vmc} and Green's function Monte Carlo (GFMC) methods \cite{gfma,gfmb}. Most of these approaches has focused on three- or four-body problems. \par\indent To treat an $N$-particle system, one needs to cope with a large number of variables required to specify the wave function. By using $(N\!\!-\!\!1)$ relative coordinates to describe the system, for example, the discretization on a mesh with $p$ points, or the expansion of the function of the relative motion between the particles in terms of $p$ suitably chosen functions leads to $p^{(N-1)}$ mesh points or basis functions, which becomes prohibitively large with increasing $N$. All but the Monte Carlo methods face this difficulty as the number of particles increases. The VMC and GFMC methods have proved to be most successful by being able to go beyond the four-nucleon problem \cite{prl95}. The secret of the efficiency of the Monte Carlo methods is the use of an importance sampling of the most relevant parts of the configuration space. This fact naturally raises a question: Even if the wave function of the $N$-particle system is expanded into an (excessively) large number of basis functions, can't one reduce the problem to a tractable one by selecting ``the most important'' basis functions ? \par\indent The aim of this paper is to present an alternative variational approach, the stochastic variational method (SVM) \cite{SVM,VSL}, by using the correlated Gaussians as basis functions \cite{corrgauss,temper}. Examples whose solutions were known before are used to demonstrate the performance of the method in treating nuclear as well as Coulomb interactions. To highlight some new physics, we have also included problems that have been hitherto unsolved. We give the formulation and some details of the method of the calculation and show applications to $N=2-7$-particle systems. \par\indent The variational foundation for the time-independent Schr\"odinger equation provides a solid and arbitrarily improvable framework for the solution of bound-state problems. The crucial point of the variational approach is the choice of the trial function. There are two widely applied strategies: (1) to select the most appropriate functional form to describe the short-range as well as long-range correlations and to compute the matrix elements by Monte Carlo technique, or (2) to use a number, possibly a great number, of simple terms, which facilitate the analytical calculation of the matrix elements. We follow the second course by using an expansion over a correlated Gaussian ``basis". \par\indent To solve the $N$-particle problem, it is of prime importance to describe the correlation between the particles properly. The correlation is conveniently represented by a correlation factor, $F=\prod_{i<j}^{N}f_{ij}$ \cite{atms,hypers,mcv,vmc,etbm}. Most calculations have used this form of $F$ directly to evaluate the matrix elements. Such calculations are, however, fairly involved beyond the three-particle system and performed by Monte Carlo integrations. An alternative way to incorporate the correlation is to approximate $f_{ij}$ as a linear combination of Gaussians ${\rm exp}(-\alpha_{ij} ({\bf r}_i-{\bf r}_j)^2)$. The $N$-particle basis function then contains product of these Gaussians: $\prod_{i<j}^{N}{\rm exp}(-\alpha_{ij}({\bf r}_i-{\bf r}_j)^2)= {\rm exp}(-\sum_{i<j}^{N}\alpha_{ij}({\bf r}_i-{\bf r}_j)^2)$. These Gaussian functions are widely used in variational calculations (see, for example, \cite{temper,etbm}). We will apply a more general form of the correlated Gaussian functions which allow for nonzero orbital angular momentum, and will use the more convenient Jacobi relative coordinates instead of the relative distance vectors. The correlated Gaussians have an important advantage. Their Hamiltonian matrix elements can be analytically calculated in a unified framework, thus enabling one to avoid the formidable calculation involving the correlation factor $F$. \par\indent The variational approximation, however, may run into difficulties for the following reasons: (i) if the nonlinear parameters specifying the basis functions are varied, it is difficult to optimize them, (ii) if they are not, then the number of terms required may be excessively large, and, in both cases, (iii) the trial function of proper symmetry becomes extremely involved. For example, conventional methods \cite{crcbgv,temper} for the choice of the Gaussian parameters lead to prohibitively large bases for more than 3 or 4 particles, which has limited the applicability of the Gaussian basis to few-body problems. \par\indent One can circumvent the optimization problem including large number of nonlinear parameters or the diagonalization of huge matrices by using the SVM. The SVM attempts to set up the most appropriate basis functions by the following stepwise procedure: One generates a would-be basis function by choosing the nonlinear parameters randomly, judges its utility by the energy gained by including it in the basis, and either keeps or discards it. One repeats this ``trial and error'' procedure until the basis set up leads to convergence. The original procedure of the SVM, proposed in \cite{SVM}, has recently been developed further and successfully applied to multicluster descriptions of light exotic nuclei, such as $^6$He=$\alpha+ n + n$, $^8$He=$\alpha+ n + n + n + n$, $^9$Li=$\alpha + t + n + n$, and $^9$C=$\alpha+^3$He$+p + p$ \cite{VSL,SVAO}. Learning from these applications, we have now generalized and refined the method further to encompass diverse systems emerging in nuclear and atomic physics. \par\indent Besides the large number of nonlinear parameters, the treatment of the increasing number of partial waves in the expansion of the wave function would also pose a formidable task. We propose here an alternative formulation to cope with this problem. Instead of using the partial wave expansion, the angular dependence of the wave function is represented by a single solid spherical harmonics whose argument contains additional variational parameters. This form makes the calculation of the matrix elements for nonzero orbital angular momentum much simpler than other methods. \par\indent It will be demonstrated that the present method has several unique features: It is based on a fully analytical calculation for most types of interactions and thus ensures high accuracy and speed. Its calculational scheme is quite universal and needs no change depending on whether the system contains nuclear or Coulombic or other interactions. It has no difficulty in treating the system of particles of unequal masses. More importantly, the wave function is obtained in a compact, analytical form and thereby can be readily used in calculations of physical properties. \par\indent As you will see later, the present method has turned out to be very accurate, and we think it is worth while to make the method and the results easily available and reproducible for interested readers. We collect all the needed ingredients of our method in order. Some of the formulae are our original developments or generalizations of known relations to $N$-particle matrix elements, and some others are collected here to make the paper self-contained. The calculation of the matrix elements presented here is different from the one of Refs. \cite{KA91,epep} in many aspects: The motion of the centre-of-mass is removed from both the Hamiltonian and the wave function. Two-particle potential matrix elements of arbitrary radial form factor are evaluated in a unified way by reducing them to the calculation of appropriate correlation functions corresponding to the interaction. The calculation of the matrix elements is extended to nonzero orbital angular momentum as well. The symmetrization postulate is imposed on the wave function at the single-particle level, which provides several advantages, especially in evaluating the matrix elements of state-dependent realistic nuclear interactions. \par\indent The organization of the paper is as follows. Section II defines the correlated Gaussian basis functions and gives the details of the stochastic procedure of selecting the basis set. Section III contains the method of calculating the matrix elements. The main steps are the calculation of matrix elements in Slater determinants (or permanents for bosons) consisting of single-particle Gaussian wave packets, the elimination of the center-of-mass motion with a very simple manipulation, and the transformation to the correlated Gaussian basis. This section also presents the modifications needed for treating systems of particles of unequal masses. Section IV presents numerical results for various systems of particles which interact via nuclear potentials or power-law potentials. Section V gives a brief summary. In the appendices the most important auxiliary formulae are collected to facilitate any future use of the formulation. \section{The correlated Gaussians and the stochastic variational method} \subsection{Basis functions} Since the variational method is always limited by the form chosen as a trial function, the trial function must be flexible enough to be able to describe the full variety of correlations between the nucleons, e.g., the short-range correlation due to the strong repulsive force, the $\alpha$-clustering typical in some light nuclei, or the long-range correlation at large distances in light halo nuclei. The correlation between the nucleons can be described by functions of appropriate relative coordinates. \par\indent Any square-integrable function with angular momentum $lm$ can be approximated, to any desired accuracy, by a linear combination of nodeless harmonic-oscillator functions (Gaussians) of continuous size parameter $a$: \begin{equation} \Gamma_{lm}({\bf r}) \sim {\rm e}^{-{\frac{1}{2}} a r^2} {\cal Y}_{lm}({\bf r}),\ \ \ {\rm with}\ \ \ {\cal Y}_{lm}({\bf r})=r^l Y_{lm}(\hat{\bf r}). \end{equation} A generalization of this to $N$-nucleon systems contains a product of the Gaussians as mentioned in the previous section. It is convenient to use a set of the Jacobi coordinates ${\bf x}=({\bf x}_1,...,{\bf x}_{N-1})$, instead of $N(N\!\!-\!\!1)/2$ relative distance vectors $({\bf r}_i-{\bf r}_j)$. An $N$-nucleon basis function, a so-called correlated Gaussian, then looks like \begin{equation} \psi_{(LS)JMTM_T}({\bf x}, A)={\cal A} \bigl\{ {\rm e}^{-{1 \over 2}\tilde{{\bf x}}A {\bf x}} \left[ \theta_{L}({\bf x}) \chi_{S}\right]_{JM} {\cal X}_{TM_T}\bigr\}, \label{corrGauss} \end{equation} where $\tilde{{\bf x}}$, the transpose of ${\bf x}$, stands for the row vector comprising the Jacobi coordinates. $\chi$ and ${\cal X}$ are the spin and isospin functions. $A$ is an $(N\!\!-\!\!1)\!\!\times\!\!(N\!\!-\!\!1)$ positive-definite, symmetric matrix of nonlinear parameters, specific to each basis element, and the quadratic form, $\tilde{\bf x}A{\bf x}$, involves scalar products of the Cartesian vectors: \begin{equation} \tilde{{\bf x}} A {\bf x}=\sum_{i=1}^{N-1} \sum_{j=1}^{N-1} A_{ij} {\bf x}_i\cdot{\bf x}_j . \end{equation} The operator ${\cal A}$ is an antisymmetrizer defined by \begin{equation} {\cal A}={1 \over \sqrt{N!} }\sum_P^{N!} {\rm sign}(P) P, \end{equation} where the sum runs over all permutations of the $N$ nucleon indices and sign($P$) stands for the parity of the permutation $P$. For a system of identical bosons, the antisymmetrizer is to be replaced with a symmetrizer. For a general case ${\cal A}$ is to represent the operator that imposes the proper symmetry on the wave function. \par\indent The function $\theta_{LM_L}({\bf x})$ in Eq. (2), which represents the angular part of the wave function, is a generalization of $\cal Y$ and can be chosen as a vector-coupled product of solid spherical harmonics of the Jacobi coordinates \begin{equation} \theta_{LM_L}({\bf x})=\left[[[{\cal Y}_{l_1}({\bf x}_1){\cal Y}_{l_2} ({\bf x}_2)]_{L_{12}} {\cal Y}_{l_3}({\bf x}_3)]_{L_{123}} ,...\right]_{LM_L} . \end{equation} Each relative motion has a definite angular momentum in Eq. (5). It may be important, however, to include several sets of angular momenta $(l_1,l_2,...,l_{N-1};L_{12},L_{123},...)$ for a realistic description. The various possible partial wave contributions increase the basis dimension; moreover, the calculation of matrix elements for this choice of $\theta_{LM_L}({\bf x})$ becomes too complicated. This choice is apparently inconvenient especially as the number of nucleons increases. To avoid this, we propose a different choice as the generalization of $\cal Y$: \begin{equation} \theta_{LM_L}({\bf x})=\eta_{KLM_L}({\bf u},{\bf x})=v^{2K+L} Y_{LM_L} ({\hat{\bf v}}),\ \ \ \ {\rm with}\ \ \ {\bf v}=\sum_{i=1}^{N-1} u_i {\bf x}_i. \end{equation} Only the total orbital angular momentum appears in this expression and it contains a parameter $\tilde{\bf u}=(u_1,...,u_{N-1})$. The vector {\bf u} may be considered as a variational parameter and one may try to minimize the energy functional with respect to it. It defines a linear combination of the Jacobi coordinates, ${\bf v}$, and the wave function of the system is expanded in terms of its angle $\hat{\bf v}$. The minimization amounts to finding the most suitable angle or a linear combination of angles. The factor of $v^{2K+L}$ plays an important role in improving the short-range behavior of the wave function. A remarkable advantage of this form of $\theta_{LM_L}({\bf x})$ is that the calculation of matrix elements becomes much simpler than in the former case because the coupling of $(N\!-\!1)$ angular momenta is completely avoided. \par\indent The two forms of $\theta_{LM_L}({\bf x})$ are in fact closely related to each other. Any of the functions of Eq. (5) may be expressed in terms of a linear combination of the terms, $v^{2K+L} Y_{LM_L} ({\hat{\bf v}})$, by using some appropriate sets of $\bf u$ values provided that each term satisfies the condition $2K\!+\!L{\le}l_1\!\!+\cdot\cdot\cdot+\!l_{N-1}$ and contains a monomial of degree $l_1+\cdot\cdot\cdot+l_{N-1}\!-\!2K\!-\!L$ in the variables, ${{\bf x}_1}^2,...,{{\bf x}_{N-1}}^2$. Therefore, if one can calculate the matrix elements using $\theta_{LM_L}({\bf x})$ defined in Eq. (6), then those with the previous form of $\theta_{LM_L}({\bf x})$ can be obtained readily. \par\indent The correlated Gaussian basis with the function $\theta_{LM_L}({\bf x})$ of Eq. (6) has parity $(-1)^L$. To construct a function with parity $(-1)^{L+1}$, Eq. (6) must be slightly generalized, e.g., to \begin{equation} \theta_{LM_L}({\bf x})=[\eta_{KL}({\bf u},{\bf x})\eta_{01} ({\bf u}',{\bf x})]_{LM_L}. \end{equation} \par\indent To assure positive definiteness, the matrix $A$ in Eq. (2) is in general expressed as $A =\tilde{G}A'G$, where $G$ is an $(N\!-\!1)\!\times\! (N\!-\!1)$ orthogonal matrix containing $(N\!-\!1)(N\!-\!2)/2$ parameters and $A'$ is a diagonal matrix, $(A')_{ij}=a'_{i} \delta_{ij}$, including $(N\!-\!1)$ positive parameters $a'_i$. Although no restriction on the parameters of the matrix $G$ is in principle necessary, it is advisable to avoid too many variables if possible. The most naive choice would be to take $G$ as a unit matrix, which is equivalent to using only a single set of the Jacobi coordinates, and then to try to reach convergence by including higher partial waves successively. Many examples show \cite{crcbgv,SVAO}, however, that this does not work well because the convergence is generally slow and moreover the computational cost of using high partial waves is quite expensive. \par\indent The matrix $G$ can also be chosen as one of the rotation matrices that connect the set of the Jacobi coordinates to other sets of independent relative coordinates. Figures 1a-1f show all topologically different sets of independent relative coordinates for a system of six identical particles. The set of coordinates in Fig. 1a is what we call the set of the Jacobi coordinates ${\bf x}$. A correlation conforming to a specific set of relative coordinates $\tilde{\bf x}'=({\bf x}'_1,...,{\bf x}'_{N-1})$ can be most efficiently described by tailoring the form of the basis function to this set of relative coordinates, that is, by using the form, exp$\lbrace -\frac{1}{2}\sum_{i=1}^{N-1}a'_i{\bf x}'_i\cdot{\bf x}'_i \rbrace$. Since the coordinates ${\bf x}'$ can be obtained by an appropriate rotation ${\cal R}$ of the Jacobi coordinates as ${\bf x}'= {\cal R}{\bf x}$, the basis function of such type can be clearly encompassed in the trial function of Eq. (2) by choosing $G={\cal R}$. The correlated Gaussian basis thereby allows for various correlations between the nucleons and different asymptotics at large distances flexibly. Depending on the character of the problem a more general choice of $G$ might be necessary. \par\indent By selecting a set of basis functions $\lbrace \psi_i; i=1,...,{\cal K}\rbrace$ [$\psi_i\equiv \psi_{(L_iS_i)JMTM_T}({\bf x}, A_i)$] that adequately spans the state space, the wave function of the $N$-nucleon system can be expanded as \begin{equation} \Psi=\sum_{i=1}^{\cal K} c_i \psi_i , \end{equation} where $\tilde{\bf c}=(c_1,...,c_{\cal K})$ is the set of linear variational parameters. The Ritz variational method defined by this trial function reduces to the generalized algebraic eigenvalue problem \begin{equation} {\cal H} {\bf c} = E {\cal N} {\bf c}, \end{equation} where ${\cal H}$ and ${\cal N}$ are, respectively, the matrices of the Hamiltonian and of the overlap \begin{equation} {\cal H}_{ij}=\langle \psi_i \vert H \vert \psi_j \rangle, \ \ {\rm and} \ \ \ \ {\cal N}_{ij}=\langle \psi_i \vert \psi_j \rangle \ \ \ \ (i,j=1,...,{\cal K}). \end{equation} \subsection{Stochastic selection of parameters and solution of the eigenvalue problem} Since a linear combination of the correlated Gaussians forms a dense set, there are different sets of $A$ that represent the wave function equally well. This enables one to select the most appropriate parameters randomly. We set up the basis stepwise by choosing $A$ from a preset domain of the parameter space and increase the basis dimension one by one. In the first step we select a number of parameter sets $A$ randomly, and keep the one that gives the lowest energy. Next we generate a new random set and calculate the energy with this two-element basis. As one more basis state always lowers the energy, we quantify its ``utility'' by the energy gained by including it in the basis. If the energy gain is larger than a preset value, $\epsilon$, then we admit this state to the basis, otherwise we discard it and try a new random candidate. This is repeated until the energy converges. The rate of convergence can be controlled by dynamically decreasing the value of $\epsilon$ during the search. This procedure is more advantageous than the earlier versions \cite{SVM,VSL} and, although not a full optimization, results in very good and relatively small bases. A similar procedure, called ``stochastic diagonalization'' has been used to determine the smallest eigenvalue of extremely large matrices \cite{SD}. \par\indent To have an economical algorithm for setting up the basis by a trial and error method, one has to find an efficient way to solve the eigenvalue problem, Eq. (9). The full diagonalization is rather time consuming and in fact unnecessary because (i) in the $({\cal K}\!\!+\!\!1)$th step of the procedure we can use the result of the ${\cal K}$th step and (ii) to judge the usefulness of a would-be basis state, only the lowest eigenvalue is needed. Let us assume that in the ${\cal K}$th step the Hamiltonian matrix ${\cal H}$ is diagonalized; its eigenenergies are $E_1 \le E_2 \le ... \le E_{\cal K}$ and its corresponding normalized eigenfunctions are ${\Psi_1},...,{\Psi_{\cal K}}$. The eigenvalue problem in the $({\cal K}\!\!+\!\!1)$th step takes the form \begin{eqnarray} & &\left( \begin{array}{ccccc} E_1 & 0 & ... & 0& \langle {\Psi_1} \vert H \vert \psi_{{\cal K}+1} \rangle \\ 0 & & & & \\ \vdots & & & \vdots & \vdots \\ 0 & & ... & E_{\cal K} & \langle {\Psi_{\cal K}} \vert H \vert \psi_{{\cal K}+1} \rangle \\ \langle {\psi_{{\cal K}+1}} \vert H \vert \Psi_{1} \rangle & & ... & \langle {\psi_{{\cal K}+1} }\vert H \vert { \Psi_{\cal K}}\rangle & \langle {\psi_{{\cal K}+1} }\vert H \vert { \psi_{{\cal K}+1}}\rangle \\ \end{array} \right) \left( \begin{array}{c} c_1 \\ \vdots \\ \vdots \\ c_{\cal K} \\ c_{{\cal K}+1} \\ \end{array} \right) \nonumber \\ & &=E \left( \begin{array}{ccccc} 1 &0 & ... & 0 & \langle {\Psi_1} \vert \psi_{{\cal K}+1} \rangle \\ 0 & & & & \\ \vdots & & & \vdots & \vdots \\ 0 & & ... & 1 & \langle {\Psi_{\cal K}} \vert \psi_{{\cal K}+1} \rangle \\ \langle \psi_{{\cal K}+1} \vert {\Psi_{1}} \rangle & & ... & \langle { \psi_{{\cal K}+1} } \vert \Psi_{\cal K}\rangle & \langle { \psi_{{\cal K}+1} } \vert \psi_{{\cal K}+1}\rangle \\ \end{array} \right) \left( \begin{array}{c} c_1 \\ \vdots \\ \vdots \\ c_{\cal K} \\ c_{{\cal K}+1} \end{array} \right). \end{eqnarray} By using the Gram-Schmidt orthogonalization method, that is, by defining \begin{equation} \vert {\bar\psi_{{\cal K}+1}} \rangle={\vert \psi_{{\cal K}+1}\rangle -\sum_{i=1}^{\cal K} \vert {\Psi_{i} } \rangle \langle {\Psi_{i}} \vert \psi_{{\cal K}+1} \rangle \over ( \langle \psi_{{\cal K}+1} \vert\psi_{{\cal K}+1} \rangle - \sum_{i=1}^{\cal K} \langle \psi_{{\cal K}+1} \vert{\Psi_{i}} \rangle \langle{\Psi_{i}} \vert\psi_{{\cal K}+1} \rangle )^{1/2}} , \end {equation} this generalized eigenvalue equation can be reduced to the conventional form \begin{equation} \left( \begin{array}{ccccc} E_1 & 0 & ... & 0& q_1 \\ 0 & & & & \\ \vdots & & & & \vdots \\ 0 & & ... & E_{\cal K} & q_{\cal K} \\ q_1 & & ... & q_{\cal K} & a \\ \end{array} \right) \left( \begin{array}{c} c_1 \\ \vdots \\ \vdots \\ c_{\cal K} \\ c_{{\cal K}+1} \\ \end{array} \right) =E \left( \begin{array}{c} c_1 \\ \vdots \\ \vdots \\ c_{\cal K} \\ c_{{\cal K}+1} \\ \end{array} \right) , \end{equation} where \begin{equation} q_i=\langle {\Psi_i} \vert H \vert {\bar\psi_{{\cal K}+1}} \rangle, \ \ \ \ \ \ a=\langle {\bar\psi_{{\cal K}+1}} \vert H \vert {\bar\psi_{{\cal K}+1}} \rangle . \end{equation} The eigenvalues are easily obtained by finding the roots of the secular equation \begin{equation} \lambda(E)\equiv\prod_{i=1}^{\cal K} (E_{\cal K}-E)\left((a-E)-\sum_{j=1}^{\cal K} {q_j^2\over E_j - E } \right)=0 . \end{equation} This secular equation has $({\cal K}\!\!+\!\!1)$ roots $\lbrace {E'}_i;i=1,...,{\cal K}\!\!+\!\!1\rbrace$ fulfilling the inequalities ${E'}_1\le E_1\le{E'}_2\le E_2\le ... \le E_{\cal K}\le {E'}_{{\cal K}+1}$. The eigenvectors are readily obtained after substituting the eigenvalues $\lbrace {E'}_i;i=1,...,{\cal K}\!\!+\!\!1\rbrace$ into Eq. (13). Note that one has to determine only the lowest eigenvalue ${E'}_1$ for the admittance criterion. \section{Calculation of the matrix elements} In this section we will give the details of the method of calculating the matrix elements between the basis function of Eq. (2). The calculation consists of three steps: (A)The calculation of the matrix elements between the Slater determinants of the Gaussian wave-packet single-particle functions, (B)A transformation from the single-particle coordinate representation to the relative and center-of-mass coordinate representation, (C)An integral transformation from the Gaussian wave-packet functions to the correlated Gaussian basis. A procedure similar to steps (A) and (B) was used to manipulate algebraically the antisymmetrization operation and the transformation of the coordinates for complex cluster systems \cite{suzuki}. In step (A) the Slater determinant for the $N$-nucleon wave function is constructed by distributing the nucleons at positions $({\bf s}_1,...,{\bf s}_N)$. These position vectors serve as the generator coordinates. The Slater determinant of the Gaussian wave packets is often used in nuclear theory, e.g., in cluster model \cite{suzuki,Brink,wild,supple} and fermionic or antisymmetrized molecular dynamics \cite{feldmeier,horiuchi}. The Hamiltonian matrix elements are analytically evaluated with the use of technique of the Slater determinants \cite{Lowden,Brink}, and can be expressed as a function of the generator coordinates. In step (B) the center-of-mass motion is completely separated from the intrinsic motion, and thus the trial wave function acquires the translational invariance. The separation of the center-of-mass motion is particularly simple in this formulation. In the last step (C) the matrix elements expressed in terms of the intrinsic generator coordinates are transformed to those between the correlated Gaussian basis functions with a definite angular momentum. Some of the essential parts of the calculational scheme is our original development, and some of them is a generalization of the technique used in the nuclear cluster model (see, for example, \cite{Kami}). We also show in subsection III.D those modifications which are needed to treat the system of particles of unequal masses. \subsection{Slater determinants of Gaussian wave packets} The $i$th nucleon with mass $m$, spin $\sigma_i$ and isospin $\tau_i$ is to be put in the single-particle Gaussian wave packet \begin{equation} {\hat \varphi}_{{{\bf s}_i}{\sigma_i}{\tau_i}}^{\nu}({\bf r}_i)= \varphi_{{\bf s}_i}^{\nu}({\bf r}_i) \chi_{{1\over 2}\sigma_i}{\cal X}_{{1\over 2}\tau_i}, \end{equation} with \begin{equation} \varphi_{{\bf s}_i}^{\nu}({\bf r}_i)=\left({2 \nu \over \pi} \right)^{3/4} {\rm e}^{-\nu ({\bf r}_i-{\bf s}_i)^2},\ \ \ \ \ {\rm and} \ \ \ \nu={m \omega \over 2 \hbar}, \end{equation} where ${\bf r}_i$ is the position vector of the nucleon, $\chi_{{1\over 2}\sigma_i}$ and ${\cal X}_{{1\over 2}\tau_i}$ are its spin and isospin function. The angular frequency $\omega$ is not a variational parameter and may be taken an arbitrary constant. The ${\bf s}_i$ parameter or ``generator'' coordinate will be used in an integral transformation to derive the matrix elements between the Gaussian basis functions. A Slater determinant of these Gaussian packets is defined by \begin{equation} \phi_{\kappa}({\bf s}_1,...,{\bf s}_N)={\cal A} \Biggl\{ \prod_{i=1}^{N}{\hat \varphi}_{{{\bf s}_i}{\sigma_i}{\tau_i}}^{\nu} ({\bf r}_i) \Biggr\} , \end{equation} where $\kappa=(\sigma_1\tau_1,...,\sigma_N\tau_N)$ is the set of the spin-isospin quantum numbers of the nucleons. The spins and the isospins of the nucleons are successively coupled to add up, respectively, to the total spin $S M_S$ and isospin $T M_T$ of the $N$-nucleon system: \begin{equation} {\chi}_{SM_S}= \left[[[\chi_{1\over 2}\chi_{1\over 2}]_{S_{12}} \chi_{1\over 2} ]_{S_{123}} ,...\right]_{SM_S} ,\ \ \ \ {\cal X}_{TM_T}= \left[[[{\cal X}_{1\over 2}{\cal X}_{1\over 2}]_{T_{12}} {\cal X}_{1\over 2}]_{T_{123}} ,...\right]_{TM_T} . \end{equation} To simplify the notation, the intermediate quantum numbers are suppressed in the following. The wave function in the ``generator coordinate space'' with the definite spin and isospin quantum numbers is a linear combination of the Slater determinants of the Gaussian packets: \begin{equation} \Phi_{SM_STT_z}({\bf s}_1,...,{\bf s}_N)= {\cal A} \lbrace \varphi_{{\bf s}_1}^{\nu}({\bf r}_1)... \varphi_{{\bf s}_N}^{\nu}({\bf r}_N) {\chi}_{SM_S}{\cal X}_{TM_T}\rbrace =\sum_{\kappa} c_{\kappa} \phi_{\kappa}({\bf s}_1,...,{\bf s}_N) , \end{equation} where $c_{\kappa}$ is a product of the Clebsch-Gordan coefficients needed to couple the spin and isospin as defined in Eq. (19). \par\indent The Hamiltonian of the $N$-nucleon system reads as \begin{equation} H=\sum_{i=1}^{N} {{\bf p}_i^2 \over 2 m} +\sum_{i<j}^{N} V_{ij} . \end{equation} The matrix elements of the Slater determinants can easily be evaluated using the well-known rules \cite{Lowden,Brink}. To make this paper self-contained, we have collected all the needed ingredients in Appendices A, B and C. The overlap of the Slater determinants is found to take the form \begin{equation} \langle \Phi_{SM_STM_T}({\bf s}_1,...,{\bf s}_N) \vert \Phi_{SM_STM_T}({{\bf s}'}_1,...,{{\bf s}'}_N) \rangle= \sum_{i=1}^{n_{o}} C_{i}^{(o)} {\rm e}^{-{1\over 2} \tilde{{\bf s}} A_{i}^{(o)} {\bf s}} , \end{equation} where $A_{i}^{(o)}$ is a $2N\!\times\!2N$ real, symmetric matrix and $\tilde{{\bf s}}$ stands for the $2N$-dimensional row vector comprising the single-particle generator coordinates, $({\bf s}_1,...,{\bf s}_N,{\bf s}'_1,...,{\bf s}'_N)$. To simplify the notation, we refer to the set of the vectors $({\bf s}'_1,...,{\bf s}'_N)$ alternatively as $({\bf s}_{N+1},...,{\bf s}_{2N})$. Note, therefore, that the quadratic form, $\tilde{{\bf s}} A_{i}^{(o)} {\bf s}$, reads as \begin{equation} \tilde{{\bf s}} A_{i}^{(o)} {\bf s}=\sum_{j=1}^{2N} \sum_{k=1}^{2N} \left(A_{i}^{(o)}\right)_{jk} {\bf s}_j\cdot{\bf s}_k . \end{equation} The matrix elements of the kinetic energy operator can also be expressed in terms of the same $C_{i}^{(o)}$'s and $A_{i}^{(o)}$'s as \begin{eqnarray} & & \langle \Phi_{SM_STM_T}({\bf s}_1,...,{\bf s}_N) \vert \sum_{i=1}^{N} {{\bf p}_i^2 \over 2 m} \vert \Phi_{SM_STM_T}({{\bf s}'}_1,...,{{\bf s}'}_N) \rangle \nonumber \\ & &={\hbar \omega \over 2} \sum_{i}^{n_o} C_{i}^{(o)} \left({3 \over 2}N-{1\over 2} \tilde{{\bf s}} A_{i}^{(o)} {\bf s} \right) {\rm e}^{-{1\over 2} \tilde{{\bf s}} A_{i}^{(o)} {\bf s}}. \end{eqnarray} The matrix elements of any term of the two-body interaction can be expressed as an integral of the two-particle correlation function multiplied by the radial form factor, $V(r)$, of the term $V_{ij}$ as below: \begin{eqnarray} & &\langle \Phi_{SM_STM_T}({\bf s}_1,...,{\bf s}_N) \vert \sum_{i<j}^{N} V_{ij} \vert \Phi_{S'M_S'T'M_T'}({{\bf s}'}_1,...,{{\bf s}'}_N) \rangle \nonumber \\ & & =\int d{\bf r} V(r) {\rm e}^{-\nu r^2} \sum_{i=1}^{n_{p}} C_{i}^{(p)} P_{i}({\bf s},{\bf r}) {\rm e}^{-{1\over 2} {\tilde{\bf s}} A_{i}^{(p)} {\bf s}+{\bf d}_i \cdot {\bf r}}, \end{eqnarray} where $P_{i}({\bf s},{\bf r})$ is a polynomial of ${\bf s}$ and ${\bf r}$, $A_{i}^{(p)}$ a $2N\!\times\!2N$ symmetric matrix, and ${\bf d}_i\cdot{\bf r}$ takes the form \begin{equation} {\bf d}_i\cdot{\bf r}=\sum_{j=1}^{2N} {\cal D}_{(i)j}{\bf s}_j \cdot{\bf r}. \end{equation} The polynomial part reduces to unity ($P_{i}({\bf s},{\bf r})=1$) in the case of pure central forces, but it has a rather simple form for spin-orbit, tensor, and other interactions as well. See Appendices B and C. The $C_{i}^{(o)}$'s, $A_{i}^{(o)}$'s, $C_{i}^{(p)}$'s, and $A_{i}^{(p)}$'s, etc. are obtained with the use of mathematical manipulation languages or fortran programs. See Appendix C. \subsection{Transformation to relative and center-of-mass coordinates} To eliminate the center-of-mass motion, we transform the single-particle coordinates to the relative and center-of-mass coordinates. For this purpose we choose one particular set of relative coordinates, the Jacobi coordinates, which is expressed in terms of the single-particle coordinates ${\bf r}_i$ as \begin{equation} {\bf x}_i=\sum_{k=1}^{N} U_{ik} {\bf r}_k \ \ \ \ (i=1,...,N), \end{equation} where the transformation matrix $U$ is defined by \begin{equation} U= \left( \begin{array}{ccccc} -1 & 1 & 0 &... & 0 \\ -{1 \over 2} & -{1 \over 2} & 1 &... & 0 \\ \vdots & & & &\vdots \\ -{1 \over N-1}&-{1\over N-1} & ... &... & 1 \\ {1\over N} &{1\over N} & ... &... &{1 \over N} \end{array} \right), \ \ \ \ \ \ {\rm and}\ \ \ \ \ U^{-1}= \left( \begin{array}{ccccc} -{1 \over 2} & -{1 \over 3} &... & -{1 \over N} & 1 \\ {1 \over 2} & -{1 \over 3} & & -{1 \over N} & 1 \\ 0 & {2 \over 3} & & \vdots & \vdots \\ \vdots & \vdots & & \vdots & \vdots \\ 0 & 0 & ... & {N-1 \over N} & 1 \end{array} \right). \end{equation} Similarly, the single-particle generator coordinates are transformed to the relative and center-of-mass generator coordinates: \begin{equation} {\bf S}_i=\sum_{k=1}^{N} U_{ik} {\bf s}_k, \ \ \ \ \ \ {\bf S}_i'=\sum_{k=1}^{N} U_{ik} {\bf s}_k' \ \ \ \ \ \ (i=1,...,N). \end{equation} The reduced masses corresponding to the transformation $U$ are given by \begin{equation} \mu_i={i \over i+1} m \ \ \ \ (i=1,...,N\!\!-\!\!1), \ \ \ \ \ \ \ {\rm and} \ \ \ \ \ \mu_N=N m. \end{equation} The product of the Gaussian single-particle wave packets can then be written as a product of Gaussians depending on the relative and center-of-mass coordinates: \begin{equation} \prod_{i=1}^{N} \varphi_{{\bf s}_i}^{\nu}({\bf r}_i)= \prod_{i=1}^{N} \varphi_{{\bf S}_i}^{\gamma_i}({\bf x}_i) , \end{equation} with \begin{equation} \gamma_i={\mu_i \omega \over 2 \hbar}. \end{equation} By using Eq. (31) and noting that the last factor of the product depends only on the center-of-mass coordinate, which is symmetric under the exchange of nucleons, the $N$-nucleon wave function can be rewritten as \begin{equation} \Phi_{SM_STM_T}({\bf s}_1,...,{\bf s}_N)= \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1}) \varphi_{{\bf S}_N}^{\gamma_N}({\bf x}_N) , \end{equation} which defines the intrinsic function that depends solely on the relative coordinates, \begin{equation} \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1})= {\cal A} \lbrace \varphi_{{\bf S}_1}^{\gamma_1}({\bf x}_1) ,..., \varphi_{{\bf S}_{N-1}}^{\gamma_{N-1}}({\bf x}_{N-1}) {\chi}_{SM_S}{\cal X}_{TM_T}\rbrace . \end{equation} As will be shown in the next subsection, this function serves as a generating function of the correlated Gaussian basis. \par\indent The Hamiltonian of Eq. (21) can be recast to the relative plus center-of-mass terms as \begin{equation} H=\left( \sum_{i=1}^{N-1} {{{\bf P}_i}^2 \over 2 \mu_i}+ \sum_{i<j}^{N} V_{ij} \right)+{{{\bf P}_N}^2 \over 2 \mu_N}\ \ {\equiv}\ \ H_{rel}+T_{cm} , \end{equation} where ${\bf P}_i$ is the momentum canonically conjugate to the Jacobi coordinate ${\bf x}_i$. The matrix elements of the Hamiltonian are then \begin{eqnarray} & & \langle \Phi_{SM_STM_T}({\bf s}_1,...,{\bf s}_N) \vert H \vert \Phi_{S'M_S'T'M_T'} ({{\bf s}'}_1,...,{{\bf s}'}_N) \rangle \nonumber \\ & &=\langle \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1}) \vert H_{rel} \vert \Psi_{S'M_S'T'M_T'}({{\bf S}'}_1,...,{{\bf S}'}_{N-1}) \rangle \langle\varphi_{{\bf S}_N}^{\gamma_N} \vert \varphi_{{{\bf S}'}_N}^{\gamma_N}\rangle \nonumber\\ & & +\langle \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1}) \vert \Psi_{S'M_S'T'M_T'}({{\bf S}'}_1,...,{{\bf S}'}_{N-1}) \rangle \langle\varphi_{{\bf S}_N}^{\gamma_N} \vert T_{cm} \vert \varphi_{{{\bf S}'}_N}^{\gamma_N}\rangle . \end{eqnarray} The matrix elements of the Hamiltonian $H_{rel}$ can be expressed by integrating over ${\bf S}_N$ \begin{eqnarray} & &\langle \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1}) \vert H_{rel} \vert \Psi_{S'M_S'T'M_T'}({{\bf S}'}_1,...,{{\bf S}'}_{N-1}) \rangle \nonumber \\ & & = \left({{\gamma_N} \over 2\pi}\right)^{3/2} \int d{\bf S}_N \langle \Phi_{SM_STM_T}({\bf s}_1,...,{\bf s}_N) \vert H \vert \Phi_{S'M_S'T'M_T'} ({{\bf s}'}_1,...,{{\bf s}'}_N) \rangle , \end{eqnarray} where use has been made of the single-particle matrix elements in Appendix A and the formula \begin{equation} \int d{\bf S}_N \left(-{{\gamma_N} \over 2}({\bf S}_N -{{\bf S}'}_N)^2 \right)^n {\rm e}^{-{{\gamma_N} \over 2}({\bf S}_N -{{\bf S}'}_N)^2}= \left(-{1\over 2}\right)^{n}(2n+1)!! \left({2\pi \over {\gamma_N}}\right)^{3\over2}. \end{equation} \par\indent Equation (37) shows that the matrix elements between the intrinsic function can be obtained in a simple way by factorizing the ${\bf S}_N$-dependent terms from the matrix elements in the single-particle basis: The quadratic forms in Eqs. (22), (24) and (25), with the help of Eq. (36), should take the form \begin{equation} \tilde{{\bf s}} A_i^{(k)} {\bf s}=\tilde{{\bf S}} B_i^{(k)} {\bf S}+ {\gamma_N} ({\bf S}_N-{{\bf S}'}_N)^2 \ \ \ \ (k=o,p). \end{equation} Here the $(2N-2)$-element column vector ${\bf S}$ is an abbreviation for the set of the Jacobi generator-coordinate vectors $({\bf S}_1,...,{\bf S}_{N-1},{{\bf S}'}_1,...,{{\bf S}'}_{N-1})$. The matrix $B_i^{(k)}$ is the $(2N\!\!-\!\!2)\!\!\times\!\!(2N\!\!-\!\!2)$ symmetric matrix defined by dropping the $N$th and $2N$th rows and columns of the matrix $\tilde{T} A_i^{(k)} T$, where \begin{equation} T= \left( \begin{array}{cc} U^{-1} & 0 \\ 0 & U^{-1} \end{array} \right) . \end{equation} It is clear from Eq. (36) that the polynomials $P_{i}({\bf s},{\bf r})$ and the vectors ${\bf d}_i$ defined in Eq. (25) can depend only on the relative generator coordinates. The dependence of the matrix elements on the center-of-mass variables, ${\bf S}_N$ and ${\bf S}'_N$, can, therefore, be factorized and the integration in Eq. (37) reduces to such a simple form as the one in Eq. (38). After the integration over ${\bf S}_N$ the matrix elements can be expressed in terms of the Jacobi generator coordinates in the form: \begin{equation} \langle \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1}) \vert \Psi_{SM_STM_T}({{\bf S}'}_1,...,{{\bf S}'}_{N-1}) \rangle= \sum_{i=1}^{n_{o}} C_{i}^{(o)} {\rm e}^{-{1\over 2} \tilde{\bf S} B_{i}^{(o)} {\bf S}}, \end{equation} \begin{eqnarray} & & \langle \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1}) \vert \sum_{i=1}^{N} {{\bf P}_i^2 \over 2 \mu_i} \vert \Psi_{SM_STM_T}({{\bf S}'}_1,...,{{\bf S}'}_{N-1}) \rangle \nonumber \\ & &={\hbar \omega \over 2} \sum_{i}^{n_o} C_{i}^{(o)} \left({3 \over 2}(N-1)-{1\over 2} \tilde{\bf S} B_{i}^{(o)} {\bf S}\right) {\rm e}^{-{1\over 2} \tilde{\bf S} B_{i}^{(o)} {\bf S}}, \end{eqnarray} \begin{eqnarray} & &\langle \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1}) \vert \sum_{i<j}^{N} V_{ij} \vert \Psi_{S'M_S'T'M_T'}({{\bf S}'}_1,...,{{\bf S}'}_{N-1}) \rangle \nonumber \\ & & =\int d{\bf r} V(r) {\rm e}^{-\nu r^2} \sum_{i=1}^{n_{p}} C_{i}^{(p)} P_{i}({\bf S},{\bf r}) {\rm e}^{-{1\over 2} \tilde{\bf S} B_{i}^{(p)} {\bf S}+{\bf D}_i \cdot {\bf r}}. \end{eqnarray} Here the convention of renumbering the set of the vectors $({{\bf S}'}_1,...,{{\bf S}'}_{N-1})$ of the ket as $({\bf S}_N,..., {\bf S}_{2N-2})$ is used to simplify the notation and thus ${\bf S}$ stands for the set of the vectors $({\bf S}_1,...,{\bf S}_{2N-2})$. The vector ${\bf D}_i$ is given by $\sum_{j=1}^{2N-2} {\hat{D}}_{(i)j} {\bf S}_j $, where the ${\hat D}_{(i)j}$'s are formed from the elements of the row vector $\tilde{\cal D}_{(i)}T$ by omitting its $N$th and $2N$th columns. The column vector ${\cal D}_{(i)}$ is defined in Eq. (26). \subsection{Integral transformation to the correlated Gaussian basis} \par\indent Here we show how to evaluate the matrix elements in the correlated Gaussian basis. Let us choose the correlated Gaussians with the form of Eq. (6) and introduce the function \begin{equation} f_{KLM}({\bf u},{\bf x},A)=\eta_{KLM} ({\bf u},{\bf x}){\rm e}^{-{1\over 2} \tilde{\bf x} A {\bf x}}, \end{equation} where \begin{equation} \eta_{KLM}({\bf u},{\bf x})=v^{2K+L} Y_{LM}({\hat{\bf v}}),\ \ \ \ {\rm with} \ \ \ \ \ {\bf v}=\sum_{i=1}^{N-1} u_i {\bf x}_i=\tilde{\bf u}{\bf x}, \end{equation} and where $\tilde{\bf u}=(u_1,...,u_{N-1})$. Note that $\bf u$ is a set of $(N\!-\!1)$ real numbers, whereas $\bf x$ are the $(N\!-\!1)$ Jacobi coordinates. The calculation of the matrix elements becomes simple if one uses a generating function of the correlated Gaussian. In fact, the following function $g$ is found to be most convenient to generate the function $f$: \begin{equation} f_{KLM}({\bf u},{\bf x},A)={1\over B_{KL}} \int d{\hat {\bf t}} Y_{LM}({\hat {\bf t}}) \left( {d^{2K+L} \over d\alpha^{2K+L}} g(\alpha,{\bf t};{\bf u},{\bf x},A) \right)_{\alpha=0 \atop t=|{\bf t}|=1}, \end{equation} where \begin{equation} g(\alpha,{\bf t};{\bf u},{\bf x},A)= {\rm e}^{-{1\over 2}\tilde{\bf x} A {\bf x} +\alpha {\bf v}\cdot{\bf t}}, \end{equation} \begin{equation} B_{nl}={4 \pi (2n+l)! \over 2^n n! (2n+2l+1)!!}. \end{equation} Equation (46) is easily proved by using the simple formula \begin{equation} ({\bf v}\cdot{\bf t})^{k}= v^k t^k \sum_{n,l \ge 0 \atop 2n+l=k} B_{nl} \sum_{m=-l}^{l} Y_{lm}({\hat {\bf v}}) Y_{lm}({\hat {\bf t}})^{\ast}. \end{equation} For a case where the function $\theta_{LM_L}({\bf x})$ of Eq. (7) is needed, the generating function $g$ of Eq. (47) must be generalized to include another factor $\alpha'{\bf v}'\cdot{\bf t}'$ . Since the following derivation remains essentially unaffected by this generalization, we will assume Eq. (6) as $\theta_{LM_L}({\bf x})$. \par\indent The generating function $g$ can be related to the product of the Gaussians centered around $\{{\bf S}_i;i=1,...,N\!-\!1\}$ through an integral transformation. To show this, we express the product of the Gaussian wave packets as \begin{equation} \prod_{i=1}^{N-1} \varphi_{{\bf S}_i}^{\gamma_i}({\bf x}_i)=\left(\frac{{\rm det}\Gamma}{\pi^ {N-1}}\right)^{3/4}{\rm e}^{-\frac{1}{2}\tilde{\bf x}\Gamma{\bf x}+ \tilde{\bf x} \Gamma{{\bf S}_H}-\frac{1}{2}{{\tilde{\bf S}}_H} \Gamma{{\bf S}_H}} \end{equation} with an $(N\!\!-\!\!1)\!\times\!(N\!\!-\!\!1)$ diagonal matrix \begin{eqnarray} \Gamma= \left( \begin{array}{cccc} 2\gamma_1 & 0 & ...& 0 \\ 0 & 2\gamma_2 & & \vdots \\ \vdots & & & \vdots \\ 0 &... & ... & 2\gamma_{N-1} \end{array} \right), \end{eqnarray} where ${{\bf S}_H}$ stands for the set of the generator coordinate vectors $({\bf S}_1,...,{\bf S}_{N-1})$. By using the familiar formula of the $n$-dimensional Gaussian integration \begin{equation} \int d{\bf x} {\rm e}^{-{1\over 2}\tilde{\bf x}A {\bf x}+\tilde{\bf T} {\bf x}}= \left({(2\pi)^{n} \over {\rm det} A }\right)^{3/2} {\rm e}^{{1\over 2}\tilde{\bf T} A^{-1} {\bf T}}, \end{equation} it is easy to prove the following equation by a direct calculation \begin{equation} g(\alpha,{\bf t};{\bf u},{\bf x},A)= \left(\frac{({\rm det}\Gamma)^{3/2}} {(4\pi)^{(N-1)/2}{\rm det}(\Gamma-A)} \right)^{3/2} {\rm e}^{-{1\over 2} \tilde{\bf T} C {\bf T}} \int d{{\bf S}_H} {\rm e}^{-{1\over 2} {{\tilde{\bf S}}_H} Q {{{\bf S}_H}}+\tilde{\bf T}{{\bf S}_H}} \left(\prod_{i=1}^{N-1} \varphi_{{\bf S}_i}^{\gamma_i}({\bf x}_i) \right), \end{equation} where $\tilde{\bf T}=({\bf T}_1,...,{\bf T}_{N-1})$, and \begin{equation} C=\Gamma^{-1}(\Gamma-A)\Gamma^{-1} ,\ \ \ \ \ Q=C^{-1}-\Gamma , \end{equation} \begin{equation} {\bf T}_i=\alpha {\bf t}\sum_{j=1}^{N-1}(\Gamma C)^{-1}_{ij} u_j \ \ \ \ (i=1,...,N\!\!-\!\!1). \end{equation} By combining Eqs. (46) and (53), the correlated Gaussian basis is found to be generated from the intrinsic state given in Eq. (34) by the integral transformation \begin{eqnarray} & & {\cal A} \{f_{KLM}({\bf u},{\bf x},A){\chi}_{SM_S}{\cal X}_{TM_T}\} ={1\over B_{KL} }\left(\frac{({\rm det}\Gamma)^{3/2}} {(4\pi)^{(N-1)/2}{\rm det}(\Gamma-A)} \right)^{3/2} \nonumber \\ &\times& \int d{\hat {\bf t}} Y_{LM}({\hat {\bf t}}) \left( {d^{2K+L} \over d\alpha^{2K+L}} {\rm e}^{-{1\over 2} \tilde{\bf T} C {\bf T}} \int d{{{\bf S}_H}} {\rm e}^{-{1\over 2} {{\tilde{\bf S}}_H} Q {{{\bf S}_H}}+\tilde{\bf T}{{{\bf S}_H}}} \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1}) \right)_{\alpha=0 \atop t=1} . \end{eqnarray} \par\indent The matrix elements between the correlated Gaussians are now easily obtained by the integral transformation from those expressed in terms of the relative generator coordinates ${\bf S}$. Using Eq. (56) gives a general formula to calculate a matrix element for any translation-invariant operator $\cal O$ \begin{eqnarray} & &\langle {\cal A} \{f_{KLM}({\bf u},{\bf x},A){\chi}_{SM_S}{\cal X}_{TM_T}\} \vert {\cal O} \vert {\cal A'} \{f_{K'L'M'}({\bf u}',{\bf x},A') {\chi}_{S'M'_S}{\cal X}_{T'M'_T}\}\rangle \nonumber \\ &=&\frac{1}{B_{KL}B'_{K'L'}}\left(\frac{({\rm det}\Gamma)^3} {(4\pi)^{(N-1)}{\rm det}(\Gamma-A){\rm det}(\Gamma-A')} \right)^{3/2} \nonumber \\ &\times& \int\!\int d{\hat {\bf t}}d{\hat {\bf t}}' Y_{LM}({\hat {\bf t}})^{*} Y_{L'M'}({\hat {\bf t}}') \left( {d^{2K+L+2K'+L'} \over d\alpha^{2K+L}d\alpha'^{2K'+L'}} {\rm e}^{-{1\over 2} \tilde{\bf T} C {\bf T}} \right.\\ &\times& \left. \int d{{\bf S}} {\rm e}^{-{1\over 2} \tilde{{\bf S}} Q {{\bf S}}+\tilde{\bf T}{{\bf S}}} \langle \Psi_{SM_STM_T}({\bf S}_1,...,{\bf S}_{N-1}) \vert {\cal O} \vert \Psi_{S'M'_ST'M'_T}({{\bf S}'}_1,...,{{\bf S}'}_{N-1}) \rangle \right)_{\alpha=\alpha'=0 \atop t=t'=1}, \nonumber \end{eqnarray} where $\tilde{\bf S}=({\bf S}_1,...,{\bf S}_{N-1},{{\bf S}'}_1={\bf S}_N,..., {{\bf S}'}_{N-1}={\bf S}_{2N-2})$ and the matrices, $C$ and $Q$, and the vectors ${\bf T}$ in Eq. (57), although the same notation is used as in Eqs. (54) and (55), are extended to include the corresponding primed quantities of the ket, that is, \begin{equation} C \longrightarrow \pmatrix{ C & 0 \cr 0 & C' \cr }, \ \ \ \ Q \longrightarrow \pmatrix{ Q & 0 \cr 0 & Q' \cr }, \ \ \ \ {\bf T} \longrightarrow \pmatrix{ {\bf T} \cr {\bf T}' \cr }. \end{equation} As is shown in Eqs. (41)-(43), the ${\bf S}$-dependence of the matrix elements is rather simple and the integration over ${\bf S}$ is done analytically. Since the variables $\alpha, \alpha', {\bf t}$, and ${\bf t}'$ appear only through the vector ${\bf T}$, those operations with respect to them as implied in Eq. (57) are performed systematically. An illustrative example is given in Appendix D. The coupling of the orbital and spin angular momenta causes no difficulty. It is very satisfactory aesthetically that matrix elements between the basis functions with any sets of the relative coordinates can be evaluated in a unified framework without any extra transformation of the coordinates. The choice of the set of the relative coordinates amounts to the choice of the matrix $A$. \subsection{Extension to the system of particles of unequal masses} \par\indent In this subsection we remark the modifications needed to treat few-particle systems containing particles of unequal masses. As you will see, most of the formulation presented in subsections III.A$-$III.C remains unchanged. Suppose that the masses of the $N$ particles are $m_1,m_2,...,m_N$. The width parameter of the Gaussian wave packet is to be changed to \begin{equation} \nu_i={m_i\omega \over {2\hbar}} \ \ \ \ (i=1,...,N). \end{equation} The overlap and Hamiltonian matrix elements are obtained as a function of the generator coordinates, $({\bf s}_1,...,{\bf s}_N,{\bf s}'_1,..., {\bf s}'_N)$, in a form similar to the previous case. In this general case of unequal masses, the matrix $U$ in Eq. (28) which defines a set of the Jacobi coordinates must be generalized to \begin{equation} U= \left( \begin{array}{ccccc} -1 & 1 & 0 &... & 0 \\ -{m_1 \over m_{12}} & -{m_2 \over m_{12}} & 1 &... & 0 \\ \vdots & & & &\vdots \\ -{m_1 \over m_{12\cdot\cdot\cdot{N-1}}} &-{m_2 \over m_{12\cdot\cdot\cdot{N-1}}} & ... &... & 1 \\ {m_1 \over m_{12\cdot\cdot\cdot{N}}} & {m_2\over m_{12\cdot\cdot\cdot{N}}} & ... &... &{m_N \over m_{12\cdot\cdot\cdot{N}}} \end{array} \right), \end{equation} where $m_{12\cdot\cdot\cdot{i}}=m_1\!+\!m_2\!+\cdot\cdot\cdot+\!m_i$. The reduced masses corresponding to this $U$ are accordingly given, instead of Eq. (30), by \begin{equation} \mu_i={m_{i+1}m_{12\cdot\cdot\cdot{i}} \over m_{12\cdot\cdot\cdot{i+1}}} \ \ \ \ (i=1,...,N\!\!-\!\!1), \ \ \ \ \ \ \ {\rm and} \ \ \ \ \ \mu_N=m_{12\cdot\cdot\cdot{N}}. \end{equation} What is important is to realize that Eq. (31), most crucial in eliminating the center-of-mass motion, still holds even with these modifications as \begin{equation} \prod_{i=1}^{N} \varphi_{{\bf s}_i}^{\nu_i}({\bf r}_i)= \prod_{i=1}^{N} \varphi_{{\bf S}_i}^{\gamma_i}({\bf x}_i). \end{equation} It is then easy to see that the rest of all the formulae are exactly the same as the case of equal masses. We can conclude that the needed modifications noted above are rather trivial and simple but still assures the elimination of the center-of-mass motion. As a simple example of unequal masses, the system of $t+d+\mu^{-}-$molecule will be considered in subsection IV.B. \section{Numerical results} This section is devoted to present the solutions of various $N\!\!=\!\!2-7$-body problems by applying the method described above. To test the method, different potentials (Yukawa, Gauss and Coulomb) have been used for bosonic and fermionic systems. Some of the examples shown here has its own physical significance, and some other solutions may be considered as benchmark test and might be useful in comparison of various few-body methods. One can expect that, besides the VMC\cite{mcv,vmc} and GFMC\cite{gfma,gfmb}, other methods will also be extended to treat more than $N=4$-particle systems. As only a few solutions are at present available for simple potentials, the examples listed here may help to test other methods. \par\indent As was discussed in subsection II.A, there is no restriction on the choice of the orthogonal matrix $G$. We have found, however, that those special rotation matrices which connect different sets of the relative coordinates especially suitable (see also \cite{SVAO}) and will use them in what follows. This greatly helps to reduce the number of parameters of $G$. In the stochastic selection of the basis elements, these special matrices $G$ and the parameters of the diagonal matrix $A'$ are randomly chosen. The vector ${\bf u}$ in Eq. (6) is also a variational parameter. To avoid an excessively large number of variational parameters we limited ${\bf u}$-vector values to those which are needed to generate the function ${\theta}_{LM_{L}}({\bf x})$ of Eq. (5) for a given set of angular momenta. Comparison of our calculation with others confirms that these limitations have not deteriorated the accuracy of the present calculation, that is, our trial function is flexible enough. In the calculation the sets of angular momenta (i.e., the sets of these special ${\bf u}$ vectors) are also randomly chosen. The main advantage of using ${\bf u}$ lies in the simple and systematic evaluation of the matrix elements from the point of view of both analytical and numerical calculations. Further test calculations will be needed to explore the utility of ${\bf u}$ as a variational parameter. \par\indent Because the dependence of the matrix elements on the variational parameters is known as is shown in Appendix D, one can organize the numerical calculation involved in the random search economically. A change of the values of the parameters does not require a recalculation of the whole matrix element. Once they have been calculated for one set of values, to calculate them for many more requires virtually no time. The average computational time is 10 minutes for a four-body and 2 hours for a six-body calculation on the VPP500 computer of RIKEN. \subsection{Few-nucleon systems} We have performed model calculations adopting different central potentials as nucleon-nucleon interaction. Some of these model problems have already been solved to high accuracy by various methods and therefore we can directly compare the solutions. The potentials used for comparison of different methods are (i) the Malfliet-Tjon (MT-V) potential \cite{mt}, which has been most extensively used as benchmark test in few-body calculations, (ii) the Volkov ``super-soft'' core potential \cite{Volkov}, (iii) the Afnan-Tang S3 (ATS3) potential \cite{ATS3} which exhibits a strong repulsive core and incorporates a difference between the spin-singlet and spin-triplet channels, (iv) and the Minnesota potential \cite{minesota} which reproduces the most important low-energy nucleon-nucleon phase shifts. The Volkov and MT-V potentials are spin-independent, while the ATS3 and Minnesota potentials are spin-dependent. The parameters of the interactions are tabulated in Table I. We choose $\hbar^2/m=41.47$ MeV fm$^2$. The Coulomb interaction is included only in calculations with the Minnesota potential where point charges are assumed and $e^2=1.44$ MeV fm. \par\indent The spins (and isospins) of the nucleons are coupled through successive intermediate couplings. The spin couplings up to $N\!\!=\!\!7$ nucleons are tabulated in Table II. One naturally expects and test calculations show that, without spin-isospin coupling, the energy convergence is much slower. The number of spin-isospin configurations rapidly increases with $N$. In the case of $^6$He, for example, assuming $S=0$ and $T=1$, the wave function has $5\!\times\! 9=45$ spin-isospin components. The number of components becomes even higher if the interaction has non-central spin-orbit, tensor, etc. parts. The nonlinear parameters are not optimized with respect to spin-isospin components, but rather, for each trial choice of the matrix $A$, we select the spin-isospin component that gives the lowest energy. Because the matrix elements between different spin-isospin components differ only in linear factors ($c_{\kappa}$ in Eq. (20)), the calculation of the matrix elements of each spin-isospin component of the wave function requires essentially the same computational effort as that needed by the calculation for only one component. \par\indent Each calculation has been repeated several times starting from different random points to check the energy convergence. The energy as a function of the number of basis states is shown in Fig. 2 for the case of $^6$Li with the Volkov potential. The energies on different random paths, after a few initial steps, approach to each other and converge to the final solution. The energy difference between two random paths as well as the tangent of the curves give us some information on the accuracy of the method on a given size of the basis. The root mean square (rms) radius of the few-nucleon system is calculated in each step and found to be rapidly convergent to its final value. By increasing the basis size the results can be arbitrarily improved when needed. \par\indent The number of basis states required to reach energy convergence increases with the number of particles but it depends on the form of the interaction as well. This latter property is illustrated in Fig. 3 for the case of the $\alpha$-particle. The soft-core Volkov potential shows rapid convergence, while the hard-core ATS3 interaction requires more basis states to get an accurate solution. The relatively fast convergence for the MT-V potential of a strong repulsion can be explained by the simplicity -- the spin-independent nature -- of this interaction. \par\indent In the following we show tables for the ground-state energies $E$ and point matter rms radii $\langle r^2\rangle^{1/2}$. The basis dimension ${\cal K}$ of the SVM listed in the tables is such that, beyond it, the energy and the radius do not change in the digits shown. Table III shows our results (SVM), together with results of others, for the application of the spin-averaged MT-V potential \cite{mt} to $N$=2--7-nucleon systems. For three-body systems, the solution of the Faddeev equation is known to be the method of choice, but the SVM can easily yield the same accurate energy. As the MT-V potential is a preferred benchmark test of the few-body calculations, there are numerous solutions available. Table III includes a few of the most accurate results. The nice agreement for four-nucleon case corroborates that the SVM is as accurate as the direct solution of the FY equations \cite{glocka}, the method of the Amalgamation of Two-body correlations into Multiple Scattering (ATMS) \cite{atms} process or the VMC \cite{wiringa} and GFMC \cite{gfma} method. The basis used in the Coupled-Rearrangement-Channel Gaussian-basis (CRCG) variational method \cite{crcbgv} is similar to that of the SVM, but the Gaussian parameters follow geometric progressions. The fact that the basis size needed in the SVM is much smaller proves the efficiency of the selection procedure. The results of the VMC calculation for the five- and six-nucleon systems are also in good agreement with the results of SVM. The MT-V potential has no exchange term; therefore, unlike the nature, it renders the five-nucleon system bound, and the nucleus tends to collapse as the binding energy increases with the number of particles. \par\indent The next example is the Volkov potential which, due to its very soft core, is the most readily solvable case. This simple potential is widely used in model calculations for light nuclei. As one sees in Table IV, the results of SVM agree with those of other calculations, especially with the one using hyperspherical harmonics (HH) functions. The number of basis states needed to reach convergence is remarkably smaller than in the case of the MT-V potential. Without the Majorana exchange term ($M=0$) this potential also leads to a collapsing system. By setting the Majorana parameter to $M$=0.6, a commonly used value to get the correct binding energy, one may obtain more reasonable energies. The Volkov potential with $M$=0.6 does not change the energies of $N$=2$-$4 nuclei, does not bind $^5$Li in accordance with the nature, but does bind the $^6$Li ground state ($E$=$-$31.82 MeV, ${\langle r^2 \rangle}^{1/2}$ =2.69 fm). \par\indent Another potential that is often used in test calculation is the ATS3 potential. We have challenged the SVM to get solution for this case because, unlike the Volkov, this spin-dependent potential has a relatively strong repulsive core (see Table I). The solution, although on a somewhat larger basis, can easily be obtained, and it is in good agreement with those of other methods in the $N$=3 system as shown in Table V. We note, however, that the energy of the SVM is significantly lower than the ones of other methods for the $\alpha$-particle. This may be due to the strong repulsion of this potential. For example, the FY calculation \cite{glocka} agrees with ours for the MT-V potential, but shows a noticeable difference in the case of the ATS3 potential. The variational calculation \cite{etbm} using a correlation factor also misses considerable energy for the ATS3 potential. Surprisingly, inspite of its exchange part, this potential also binds the five-nucleon system and overbinds the six-nucleon systems very much. \par\indent The last example for the few-nucleon system uses the more realistic Minnesota potential\cite{minesota}, which is a central interaction of Gaussian form, containing space-, spin-, and isospin-exchange operators (see Table I). The Minnesota potential has often been used in cluster-model calculations of light nuclei. Table VI shows results with this potential, where the Coulomb interaction between protons is also included. All possible spin and isospin configurations are allowed for and all spherical harmonics that give non-negligible contribution are included in $\theta_{LM_L}({\bf x})$. Since the method has proved to be accurate and reliable for other potentials, it is justifiable to view these results as testing the interaction rather than the method, and hence the results are compared with experimental data. The energy and the radius of triton and $\alpha$-particle converge, with small bases, to realistic values. The Minnesota potential, correctly, does not bind the $N$=5 system, but it binds $^6$He and slightly overbinds $^6$Li. The radius of $^6$He is found to be much larger than that of $^4$He, consistently with the halo structure of $^6$He \cite{SVAO}. It is for the first time that the Minnesota force is tested without assuming any cluster structure or restricting the model space by any other bias. The agreement is surprisingly good not only with experiment but also with cluster-model calculations for all nuclei \cite{LKBD}. \par\indent It is interesting to note that none of these simple potentials binds the four-neutron system. The Volkov potential, for example, is so strong that it binds the singlet two-neutron system, but it does not allow the neutrons to form a four-neutron bound state due to the Pauli principle. \par\indent As an example for bosonic nuclear few-body system, we consider the case of structureless $\alpha$-particles interacting via the state-independent potential; \begin{equation} V(r)=500\,\,{\rm exp}\,[-(0.7r)^2\,]-130\,\,{\rm exp}\,[-(0.475r)^2\,] \ \ \ \ ({\rm MeV}), \end{equation} where $r$ is in fm. This potential is taken from Ali-Bodmer's $S$-state potential\cite{AB}. It has a repulsive core which is about 370 MeV high and extends up to 2 fm. The repulsive core prevents the $\alpha$-particles from collapsing. The results are compared in Table VII. Our calculation agrees with the ATMS result for the $N=$3 and 4 systems. \subsection{Coulombic systems} The results for the long range $1/r$ potential are collected in this subsection. The first example is the polyelectric system $(me^+,ne^-)$. The possibility that $m$ positrons and $n$ electrons form a bound system was originally suggested by Wheeler \cite{wheeler} and this question has been extensively studied since then. Besides the trivial and analytically solvable $m=1$ and $n=1$ case, the existence of the positronium negative ion ($m=1,n=2$) was also predicted by Wheeler \cite{wheeler}. Dozens of works have attempted to solve the $e^+,e^-,e^-$ Coulombic three-body problem, continuously refining the accuracy of the calculated binding energy \cite{epe,ho}. Despite of numerous attempts, no one has obtained bound states for the polyelectric system of more than four particles. The positronium negative ion has experimentally been observed \cite{mills}. \par\indent The binding energy of the positronium molecule $(2e^+,2e^-)$ was first calculated by Hylleraas and Ore \cite{hyll}. To date, the positronium molecule has not been directly observed, and this fact intensifies the theoretical interest to solve this Coulombic four-body problem. This molecule is short-lived because the electron and positron may annihilate. Unlike the positronium ion, the positronium molecule is neutral, and therefore the best chance to distinguish it from the positronium itself is related to their different lifetimes. The QED formulae to determine the probability of a pair annihilation in the positronium molecule through a $k$-photon process ($k=0$,1,2,...) would require a highly accurate wave function \cite{frolov95}. \par\indent In Table VIII we compare our results to the most precise calculations found in the literature. The correlated Gaussian function without the polynomial part ($K=0$ in Eq. (6)) is known to poorly represent the Coulomb cusps \cite{KA91}. To improve the cusp properties the trial function with $K=0,1,2$ polynomials has been used. \par\indent As is shown in Table VIII, our calculation reproduces the first six digits of the variational calculation of Ref.~\cite{epe} for the ground state of (2e$^+$,e$^-$), and the rms radius also agrees with it. There are two recent variational calculations \cite{epep,KA93} for the positronium molecule using the correlated Gaussian functions. In these works the nonlinear parameters were determined by optimatization. To compare our calculations to theirs directly, the value of $K=0$ was chosen and the same basis size (${\cal K}$=300) was used. Our result is slightly better than the energy obtained by them and this reinforces the reliability and powerfulness of the random selection of the nonlinear parameters. The number of nonlinear parameters of this case is ${\cal K}\!\!\times \!\!(4\!\!\times \!\!3)/2=1800$. The complete optimatization of the parameters is, of course, superior to the SVM. Test calculations show that, provided the number of parameters is low, that is, a full optimatization is feasible, the optimatization finds lower energy. But when the number of parameters becomes high, the full optimatization becomes less and less practical partly because it fails to find true minimum and partly because the computation becomes too excessive. \par\indent We found no bound states for the ($3e^+,2e^-$) and ($3e^+,3e^-$) systems. The energy of ($3e^+,3e^-$), for example, converges to the sum of the energy of a dipositronium molecule and of a positronium (0.515989 a.u.+0.25 a.u.=0.765989 a.u.). Allowing the selection of the parameters from a larger region increases the rms radius, which is typical of an unbound state. The system of a negative and a positive positronium ion thus forms no bound state but dissociates into a dipositronium molecule and a positronium. \par\indent Calculations for $L=1$ state also fails to find a bound system. This result entails that the Coulomb force cannot bind more than four particles out of identical charged fermions and their antiparticles. \par\indent To examine the role of the Pauli principle in preventing five-electron-positron system from forming bound states, we repeated the same calculation replacing the fermions by bosonic equivalents. On a different scale, these systems may be identified, e.g., as the systems of $\pi^-$ and $\pi^+$ with their strong interaction neglected \cite{boson}. Such bosons turn out to form bound states even for $N=5$. As is expected, the radius of the charged boson system decreases as the number of particles increases. It is interesting to note that the energies of bosonic and fermionic systems are equal for $N=3$ and for $N=4$. The reason is that the energy minimum belongs to the same spatial configurations, that is, to a triangular pyramid for $N=4$, for example \cite{ho86}. \par\indent In Table IX we show results for bosonic and fermionic systems with a purely attractive $Gm^2/r$ (``gravitational") interaction. An $N$-body system of identical particles bound together by attractive pair potentials always collapses in large-$N$ limit (the binding energy per particle rises with $N$ to infinity), even if the particles are fermions. Self-gravitating boson systems have recently attracted some interest \cite{selfgrav}. For these systems, both variational lower and upper bounds are available. In this case even the five-fermion system is bound. Thus the lack of bound states in five-electron-positron systems is a joint effect of the antisymmetry and of the repulsion between identical particles. \par\indent Finally, we mention an example involving an excited state. With ${\cal K}=500$, the SVM gives the energies of the ground and first excited states of the $t+d+\mu^-$ system as $-$111.3640 and $-$100.9121 a.u., which are respectively compared to $-$111.364342 and $-$100.916421 a.u. of the CRCG result \cite{dtmu} with ${\cal K}=1442$, while the configuration-space Faddeev calculation \cite{faddeev} gives $-$111.36 a.u. for the ground state but no information for the excited state. \section{Summary} \par\indent We have formulated a variational calculation for few-body systems using the stochastic variational method on the correlated Gaussian basis. We have demonstrated the versatility of the correlated Gaussians and the efficiency of the stochastic variation by various numerical examples for $N=2-7$-particle systems. All the details of both formulation and calculational procedure are included to make this paper self-contained and easily reproducible. \par\indent The comparison with other calculations has corroborated the accuracy and efficiency of the method. In none of the test cases has the present method proved to be inferior to any of the alternative methods, and yet the method does not require excessive computational effort. \par\indent The correlation between the particles plays an important role in describing the few-body system realistically. It has been taken into account in the framework of the correlated Gaussian functions. The correlated Gaussians are constructed from products of the Gaussian wave-packet single-particle functions through an integral transformation, which has enabled us to evaluate the center-of-mass motion free matrix elements analytically starting from the single-particle level. The nonlinear parameters of the correlated Gaussians have been selected by the stochastic variational method with a trial and error procedure. The success of the method using the correlated Gaussian basis is probably due to the fact that none of the Gaussians is indispensable, that is, there are different sets of the Gaussian parameters that represent the wave function equally well. \par\indent The method presented in this paper can be useful to solve few-body problems in diverse fields of physics such as description of microclusters, non-relativistic quark model, and halo nuclei. Among others, the most important application is the solution of the nuclear few-body problem, that is, a description of light nuclei by using ${\it realistic}$ nucleon-nucleon potentials. In this case one has to take into account both short-range repulsion and higher orbital angular momenta required by the non-central components. Our test examples show that the correlated Gaussian basis function might be a suitable candidate to cope with these requirements. As is explained in Appendix B, the evaluation of the matrix elements for the non-central potentials poses no serious problem, and calculations including such potentials for nuclear few-body systems are under way. \par\indent The limitations of the present method are those implied by the basis size and by the computer memory to store the matrix elements in the generator coordinate space. The limitations may become excessive as the number of particles and/or spin and isospin configurations become large. \par\indent To extend the method to nuclei of larger mass number in an approximate way, one can freeze part of the model space for a group of nucleons (cluster). One can omit, for example, some of the spin or isospin channels. It might be a good approximation to consider only those spin channels where the spins of the like nucleon pairs are coupled to zero. One can also restrict the intrinsic spatial motion of a cluster by fixing the nonlinear parameters to some appropriate values. One can introduce $N$ clusters and place the nucleons of each cluster into a common harmonic oscillator well, for example. The microscopic multicluster model is based on this approximation. The matrix elements needed in this multicluster model are given as a special case of those presented in this paper. In fact, one only needs to choose the single-particle generator coordinates, $({\bf s}_1,...,{\bf s}_A)$, such that \begin{equation} {\bf s}_{n_{i-1}+1}={\bf s}_{n_{i-1}+2}=\cdot\cdot\cdot= {\bf s}_{n_{i-1}+n_i} \ \ \ \ (i=1,...,N), \end{equation} where $n_i$ is the number of nucleons in the ${\it i}$th cluster ($n_0=0,\ \ n_1+\cdot\cdot\cdot+n_N=A$), and needs to couple appropriately the spins and isospins of the nucleons in each cluster to the spin and isospin of the cluster. The microscopic multicluster model has been successfully applied for description of the structure of light nuclei (see, for example, \cite{SVAO,Brink,wild}). \par\indent \par\indent Finally, we summarize some merits of our method in the following: \begin{description} \item{(i)} Fully analytical calculational scheme; this plays a major role in the high speed and accuracy of the calculation. \item{(ii)} Universality of the scheme. One needs to introduce no change, for example, between describing a multinucleon system and a Coulombic few-body system. It is easily adaptable to identical or non-identical particles, to fermions or bosons or mixed systems. The masses of particles may be different, yet no problem with the center-of-mass motion arises. \item{(iii)} No expansion of the interaction is needed, and thus no problems in partial-wave truncation arise. \item{(iv)} The convergence of the energy is fast. If one needs just a 2--3-digit estimate of the energy, it is enough to use a very small basis. \item{(v)} The method is also accurate for excited states, which are obtained simultaneously with the same diagonalization (provided their angular momenta and parities are the same as those of the ground state; but only such excited states may be problematic). \item{(vi)} The wave function is obtained in a compact, analytical form. It is then easy to use it in calculations of physical properties. It is ``portable'', reproducible, and easily testable. \end{description} \section*{Appendix A. Single-particle matrix elements} The aim of this appendix is to list the single-particle matrix elements between Gaussian wave packets (Eq. (17)). The Gaussian packets are generalized in this appendix to have different width parameters and the expressions are therefore slightly more general than needed in the formulae of the main text. These single-particle matrix elements are, however, required for treating particles of unequal masses, in which the width parameter $\nu$ belonging to the particle of mass $m$ is to be chosen by Eq. (59). The overlap of two Gaussian wave packets is \begin{equation} \langle \varphi_{{\bf s}_1}^{\nu_1} \vert \varphi_{{\bf s}_2}^{\nu_2} \rangle =\left({2 \sqrt{\nu_1 \nu_2} \over \nu_1 +\nu_2}\right)^{3/2} {\rm exp}\,\left(-{\nu_1 \nu_2 \over \nu_1 +\nu_2} ({{\bf s}_1}- {{\bf s}_2})^2\,\right) . \end{equation} The matrix element of the kinetic energy operator $(T=-\frac{{\hbar}^2} {2M}\Delta)$ reads as \begin{equation} \langle \varphi_{{\bf s}_1}^{\nu_1} \vert T\vert\varphi_{{\bf s}_2}^{\nu_2} \rangle ={\hbar^2 \over 2M } {2\nu_1\nu_2 \over \nu_1 +\nu_2} \left( 3-{2 \nu_1 \nu_2\over \nu_1+\nu_2} ({\bf s}_1-{\bf s}_2)^2 \right) \langle \varphi_{{\bf s}_1}^{\nu_1} \vert \varphi_{{\bf s}_2}^{\nu_2} \rangle . \end{equation} The matrix element of the square radius becomes \begin{equation} \langle \varphi_{{\bf s}_1}^{\nu_1} \vert {\bf r}_1^2 \vert \varphi_{{\bf s}_2}^{\nu_2} \rangle = {1 \over 2(\nu_1 +\nu_2)} \left( 3+{2 \over \nu_1+\nu_2} (\nu_1{\bf s}_1+\nu_2{\bf s}_2)^2 \right) \langle \varphi_{{\bf s}_1}^{\nu_1} \vert \varphi_{{\bf s}_2}^{\nu_2} \rangle . \end{equation} The two-particle matrix element of a $\delta$-function is given by \begin{eqnarray} & &\langle \varphi_{{\bf s}_1}^{\nu_1}\varphi_{{\bf s}_2}^{\nu_2} \vert \delta({\bf r}_1-{\bf r}_2-{\bf r})\vert \varphi_{{\bf s}_3}^{\nu_3}\varphi_{{\bf s}_4}^{\nu_4} \rangle = \left({(\nu_1+\nu_3)(\nu_2+\nu_4) \over (\nu_1+\nu_2+\nu_3+\nu_4) \pi }\right)^{3/2} \nonumber \\ &\times& {\rm exp}\,\Biggl(-{(\nu_1+\nu_3)(\nu_2+\nu_4) \over \nu_1+\nu_2+\nu_3+\nu_4} \Bigl({\bf r}-{\nu_1{\bf s}_1+\nu_3{\bf s}_3\over \nu_1+\nu_3} +{\nu_2{\bf s}_2+\nu_4{\bf s}_4\over \nu_2+\nu_4}\Bigr)^2\,\Biggr) \langle \varphi_{{\bf s}_1}^{\nu_1} \vert \varphi_{{\bf s}_3}^{\nu_3} \rangle \langle \varphi_{{\bf s}_2}^{\nu_2} \vert \varphi_{{\bf s}_4}^{\nu_4} \rangle . \end{eqnarray} \section*{Appendix B. Matrix elements of the two-body potentials} The scope of this appendix is the calculation of the matrix elements of the different ingredients of the two-nucleon interaction between Gaussian wave packets. Most of the widely used coordinate-space two-nucleon interactions consist of central- ($O^c_{12}$), tensor- ($O^t_{12}$), spin-orbit- ($O^b_{12}$), ${\bf L}^2$- (or ${\bf p}^2$) ($O^q_{12}$) and quadratic spin-orbit ($O^{bb}_{12}$) type potentials. (We follow the abbreviated notations -- $c,t,b,q,bb$ -- invented by Urbana-Argonne group \cite{LP81,WSA84}. The definition of these operators is given in Table X.) These potential terms might be multiplied by the \mbox{\boldmath{$\tau$}}$_1\cdot$\mbox{\boldmath{$\tau$}}$_2$ ($O^\tau_{12}$), \mbox{\boldmath{$\sigma$}}$_1\cdot$\mbox{\boldmath{$\sigma$}}$_2$ ($O^\sigma_{12}$), or \mbox{\boldmath{$\tau$}}$_1\cdot$\mbox{\boldmath{$\tau$}}$_2$ \mbox{\boldmath{$\sigma$}}$_1\cdot$\mbox{\boldmath{$\sigma$}}$_2$ ($O^{\tau \sigma}_{12}$) spin-isospin operators and then one ends up with the general form \begin{equation} V_{ij}= \sum_{p} \int d{\bf r} V^{p}(r) \delta({\bf r}_i-{\bf r}_j-{\bf r}) O_{ij}^{p} , \end{equation} where $p=c,c\tau,c\sigma,c\tau\sigma,t,t\tau,b,b\tau,q,q\tau,q\sigma, q\tau\sigma, bb,bb\tau$ is the short-hand notation to specify the component and $V^p(r)$ is the corresponding radial form factor. By using Eqs. (16), (17), (68), (69), and after some straightforward transformation, the matrix element of this interaction can be written as \begin{eqnarray} & &\langle {\hat \phi}^{\nu}_{{\bf s}_1\sigma_1\tau_1} {\hat \phi}^{\nu}_{{\bf s}_2\sigma_2\tau_2} \vert V_{12} \vert {\hat \phi}^{\nu}_{{{\bf s}_1}'\sigma_1'\tau_1'} {\hat \phi}^{\nu}_{{{\bf s}_2}'\sigma_2'\tau_2'} \rangle = \sum_{p} \int d{\bf r} V^{p}(r) f_{p}(r){\cal M}({\bf r})\nonumber \\ &\times&\langle\chi_{{1\over 2}\sigma_1}{\cal X}_{{1\over 2}\tau_1}\, \chi_{{1\over 2}\sigma_2}{\cal X}_{{1\over 2}\tau_2}\vert\, r^{2}B^{p}+\sum_{l=0}^{2}\sum_{m=-l}^{l} (-1)^{m}r^{l}Y_{lm}({\hat {\bf r}}) C^{p}_{l\,-m}\,\vert\, \chi_{{1\over 2}\sigma'_1}{\cal X}_{{1\over 2}\tau'_1}\, \chi_{{1\over 2}\sigma'_2}{\cal X}_{{1\over 2}\tau'_2}\rangle, \end{eqnarray} with \begin{equation} {\cal M}({\bf r})=\Bigl({\nu \over \pi}\Bigr)^{3/2} {\rm exp}\,\Biggl(-\nu r^2+\nu ({\bf s}_1\!\!-\!\!{\bf s}_2\!\!+\!\! {\bf s}'_1\!\!-\!\!{\bf s}'_2)\cdot{\bf r}-{\nu \over 4} ({\bf s}_1\!\!-\!\!{\bf s}_2\!\!+\!\!{\bf s}'_1\!\!-\!\!{\bf s}'_2)^2\, \Biggr)\,\langle \varphi_{{\bf s}_1}^{\nu} \vert \varphi_{{\bf s}'_1} ^{\nu} \rangle \langle \varphi_{{\bf s}_2}^{\nu} \vert \varphi_{{\bf s}'_2}^{\nu} \rangle . \end{equation} The $B^{p}$ and $C^{p}_{lm}$, independent of ${\bf r}$, are the operators in spin-isospin space and are listed in Table X for the most important terms ($c,t,b,q,bb$). The remaining terms can be easily derived by multiplying these operators by the appropriate $\mbox{\boldmath $\sigma$}$ and $\mbox{\boldmath $\tau$}$ operators. The function $f_p(r)$ has the simple form: $f_p(r)=r^{-2}$ if $p=t$ or $t\tau$, and $f_p(r)=1$ otherwise. \par\indent The calculation of matrix elements of the operators appearing in Table X in the spin part is easily done with the Clebsch-Gordan coefficient. They are given below by suppressing the spin function $\chi_{\frac{1}{2} \sigma_i}\,(m_i=\frac{1}{2} \sigma_i)$. \begin{eqnarray} & &\langle m_1\,m_2\,\vert\, {\mbox{\boldmath $\sigma$}_1}\cdot{\mbox{\boldmath $\sigma$}_2}\, \vert \,m'_1\,m'_2\,\rangle \nonumber \\ &=&3(-1)^{m_1-m'_1}\delta_{m_1+m_2,m'_1+m'_2} \langle {1 \over 2}\,m'_1\,1\,m_1\!\!-\!\!m'_1\,\vert\, {1 \over 2}\,m_1\rangle \langle {1 \over 2}\,m'_2\,1\,m_2\!\!-\!\!m'_2\,\vert\, {1 \over 2}\,m_2\rangle. \end{eqnarray} \begin{eqnarray} & &\langle m_1\,m_2\,\vert\, \left[{\mbox{\boldmath $\sigma$}_1}\times{\mbox{\boldmath $\sigma$}_2} \right]^{(2)}_{m}\, \vert \,m'_1\,m'_2\,\rangle=3\,\delta_{m,m_1+m_2-m'_1-m'_2} \nonumber \\ &\times&\langle {1 \over 2}\,m'_1\,1\,m_1\!\!-\!\!m'_1\,\vert\, {1 \over 2}\,m_1\rangle \langle {1 \over 2}\,m'_2\,1\,m_2\!\!-\!\!m'_2\,\vert\, {1 \over 2}\,m_2\rangle \langle 1\,m_1\!\!-\!\!m'_1\,1\,m_2\!\!-\!\!m'_2\,\vert\,2\,m\rangle. \end{eqnarray} \begin{eqnarray} & &\langle m_1\,m_2\,\vert\, \left[{\bf x}\times ({\mbox{\boldmath $\sigma$}_1}+{\mbox{\boldmath $\sigma$}_2}) \right]^{(1)}_{m}\, \vert\,m'_1\,m'_2\,\rangle=\sqrt{3}\, \sum_{q_1,q_2=-1} ^{1}({\bf x})_{q_1} \nonumber \\ &\times&\langle 1\,q_1\,1\,q_2\,\vert\,1\,m\rangle \Bigl\{\delta_{m_2,m'_2} \langle{1\over 2}\,m'_1\,1\,q_2\,\vert\,{1\over 2}\,m_1\rangle+ \delta_{m_1,m'_1} \langle{1\over 2}\,m'_2\,1\,q_2\,\vert\,{1\over 2}\,m_2\rangle\Bigr\}. \end{eqnarray} \begin{eqnarray} & &\langle m_1\,m_2\,\vert\, ({\bf x}\cdot{\mbox{\boldmath $\sigma$}_2}) {\mbox{\boldmath $\sigma$}_{1_{m}}}+ ({\bf x}\cdot{\mbox{\boldmath $\sigma$}_1}) {\mbox{\boldmath $\sigma$}_{2_{m}}} \, \vert \,m'_1\,m'_2\,\rangle \nonumber \\ &=&3\,\Bigl\{(-1)^{m_2-m'_2}\,\delta_{m,m_1-m'_1}({\bf x})_{m'_2-m_2}+ (-1)^{m_1-m'_1}\,\delta_{m,m_2-m'_2}({\bf x})_{m'_1-m_1}\Bigr\} \\ &\times&\langle {1 \over 2}\,m'_1\,1\,m_1\!\!-\!\!m'_1\,\vert\, {1 \over 2}\,m_1\rangle \langle {1 \over 2}\,m'_2\,1\,m_2\!\!-\!\!m'_2\,\vert\, {1 \over 2}\,m_2\rangle . \nonumber \end{eqnarray} \begin{eqnarray} & &\langle m_1\,m_2\,\vert\, ({\bf x}\cdot{\mbox{\boldmath $\sigma$}_1}) ({\bf x}\cdot{\mbox{\boldmath $\sigma$}_2}) \, \vert \,m'_1\,m'_2 \rangle \, =3\,(-1)^{m_1-m'_1+m_2-m'_2} \nonumber \\ &\times&({\bf x})_{m'_1-m_1}({\bf x})_{m'_2-m_2} \langle {1 \over 2}\,m'_1\,1\,m_1\!\!-\!\!m'_1\,\vert\, {1 \over 2}\,m_1\rangle \langle {1 \over 2}\,m'_2\,1\,m_2\!\!-\!\!m'_2\,\vert\, {1 \over 2}\,m_2\rangle . \end{eqnarray} \begin{eqnarray} & &\langle m_1\,m_2\,\vert\, \left[\,\left[{\bf x}\times{\mbox{\boldmath $\sigma$}_1}\right]^{(1)} \times\left[{\bf x}\times{\mbox{\boldmath $\sigma$}_2}\right]^{(1)} \,\right]^{(2)}_{m} \, \vert \,m'_1\,m'_2\,\rangle \nonumber \\ &=&3\,\langle {1 \over 2}\,m'_1\,1\,m_1\!\!-\!\!m'_1\,\vert\, {1 \over 2}\,m_1\rangle \langle {1 \over 2}\,m'_2\,1\,m_2\!\!-\!\!m'_2\,\vert\, {1 \over 2}\,m_2\rangle\,\nonumber \\ &\times& \sum_{q_1,q_2=-1}^{1}({\bf x})_{q_1}({\bf x})_{q_2} \langle 1\,q_1\!\!+\!\!m_1\!\!-\!\!m'_1\,1\,q_2\!\!+\!\!m_2\!\!-\!\! m'_2\,\vert\,2\,m\rangle \\ &\times&\langle 1\,q_1\,1\,m_1\!\!-\!\!m'_1\,\vert\,1\,q_1\!\!+\!\!m_1\!\!-\!\!m'_1 \rangle \langle 1\,q_2\,1\,m_2\!\!-\!\!m'_2\,\vert\,1\,q_2\!\!+\!\!m_2\!\!-\!\!m'_2 \rangle. \nonumber \end{eqnarray} Here ${\bf x}$ is a 3-dimensional vector and $({\bf x})_m$ stands for its spherical component $\sqrt{4\pi\over 3} x Y_{1m}({\hat{\bf x}})$. \section*{Appendix C. Matrix elements of Slater determinants} In this appendix we briefly outline the calculation of the matrix elements between Slater determinants of Eq. (18) and show their concrete functional form, that is, the dependence on the generator coordinates ${\bf s}$. We assume that the width parameter of the Gaussian wave packet is chosen to be the same for all nucleons. The overlap of two Slater determinants is equal to the determinant of the matrix of the single-particle overlaps: \begin{equation} \langle \phi_{\kappa} ({\bf s}_1,...,{\bf s}_N) \vert \phi_{\kappa'}({{\bf s}'}_1,...,{{\bf s}'_N}) \rangle ={\rm det} \lbrace B \rbrace, \end{equation} where \begin{equation} B_{ij}= \langle {\hat\varphi}_{{\bf s}_i\sigma_i\tau_i}^{\nu} \vert {\hat\varphi}_{{{\bf s}'}_j{\sigma}'_j{\tau}'_j}^{\nu} \rangle \ \ \ \ \ \ \ (i,j=1,...,N). \end{equation} By using the definition of the determinant, this can be rewritten as \begin{equation} \langle \phi_{\kappa} ({\bf s}_1,...,{\bf s}_N) \vert \phi_{\kappa'}({{\bf s}'}_1,...,{{\bf s}'_N}) \rangle = \sum_{P}^{N!} {\rm sign}(P) \langle {\hat \varphi}_{{\bf s}_1\sigma_1\tau_1}^{\nu} \vert {\hat\varphi}_{{{\bf s}'}_{p_{1}}{\sigma}'_{p_1}{\tau}'_{p_1}}^{\nu} \rangle... \langle {\hat\varphi}_{{\bf s}_N\sigma_N\tau_N}^{\nu} \vert {\hat\varphi}_{{{\bf s}'}_{p_{N}}{\sigma}'_{p_{N}}{\tau}'_{p_{N}}} ^{\nu} \rangle , \end{equation} where $(p_{1}\cdot \cdot \cdot p_{N})$ is the permutation $P$ of the set $(1\cdot \cdot \cdot N)$. Substituting the overlap of the single-particle overlaps of Eq. (65) into Eq. (80) yields an explicit formula for the overlap of the Slater determinants: \begin{equation} \langle \phi_{\kappa} ({\bf s}_1,...,{\bf s}_N) \vert \phi_{\kappa'}({{\bf s}'}_1,...,{{\bf s}'_N}) \rangle = \sum_{P}^{N!} C_{P} {\rm e}^{-{1\over 2} \tilde{\bf s} A_{P} {\bf s}} , \end{equation} where the matrix $A_P$ is defined by \begin{equation} (A_P)_{ij}=(A_P)_{N+i,N+j}= \nu \delta_{ij},\ \ \ \ (A_P)_{i,N+j}=(A_P)_{N+j,i}=-\nu \delta_{j,p_{i}} \ \ \ \ (1\le i,j \le N), \end{equation} and \begin{equation} C_{P}={\rm sign}(P) \delta_{{\sigma_1}{\sigma'}_{p_1}} \delta_{{\tau_1}{\tau'}_{p_1}} \cdot\cdot\cdot \delta_{{\sigma_N}{\sigma'}_{p_N}} \delta_{{\tau_N}{\tau'}_{p_N}}. \end{equation} The orthogonality of the spin-isospin functions greatly reduces the number of terms in the summation over $P$. \par\indent The matrix element of the kinetic energy operator is \begin{equation} \sum_{i=1}^{N} \langle \phi_{\kappa} ({\bf s}_1,...,{\bf s}_N) \vert T_i \vert \phi_{\kappa'}({{\bf s}'}_1,...,{{\bf s}'_N}) \rangle = \sum_{i=1}^{N} \sum_{j=1}^{N} \langle{\hat \varphi}_{{{\bf s}_i}{\sigma_i}{\tau_i}}^{\nu} \vert T \vert {\hat \varphi}_{{{{\bf s}'}_j}{\sigma'_j} {\tau'_j}}^{\nu} \rangle (-1)^{i+j} {\det} \lbrace B^{ij}\rbrace, \end{equation} where $B^{ij}$ is obtained by omitting the $i$th row and the $j$th column of the matrix $B$ defined in Eq. (79). The substitution of the single-particle overlaps and the single-particle matrix elements of the kinetic energy operator enables us to obtain \begin{equation} \sum_{i=1}^{N} \langle \phi_{\kappa} ({\bf s}_1,...,{\bf s}_N) \vert T_i \vert \phi_{\kappa'}({{\bf s}'}_1,...,{{\bf s}'_N}) \rangle = {\hbar \omega \over 2} \sum_{P}^{N!} C_{P} \left({3 \over 2}N -{1\over 2} \tilde{\bf s} A_{P} {\bf s}\right){\rm e}^{-{1\over 2} \tilde{\bf s} A_{P} {\bf s}}. \end{equation} Note that the coefficients $C_{P}$ and the matrices $A_P$ in Eqs. (81) and (85) are the same. The manipulation similar to this leads us to the matrix element of the square radius as \begin{equation} \sum_{i=1}^{N} \langle \phi_{\kappa} ({\bf s}_1,...,{\bf s}_N) \vert {\bf r}_i^2 \vert \phi_{\kappa'}({{\bf s}'}_1,...,{{\bf s}'_N}) \rangle = {1 \over 2\nu} \sum_{P}^{N!} C_{P} \left({3 \over 2}N -{1\over 2} \tilde{\bf s} A_{P} {\bf s} + \nu\tilde{\bf s}{\bf s} \right){\rm e}^{-{1\over 2} \tilde{\bf s} A_{P} {\bf s}}. \end{equation} \par\indent As is explained in Appendix B, the matrix element of any two-body interaction between the Slater determinants may be reduced to the following: \begin{equation} \sum_{i<j}^{N} \langle \phi_{\kappa} ({\bf s}_1,...,{\bf s}_N) \vert \delta({\bf r}_i-{\bf r}_j-{\bf r})O_{ij}^p \vert \phi_{\kappa'}({{\bf s}'}_1,...,{{\bf s}'_N}) \rangle. \end{equation} The matrix element of Eq. (87) is called the correlation function of type $p$ evaluated between the Slater determinants. The calculation of this matrix element can be done with the use of the basic two-body matrix elements of Appendix B and the single-particle overlaps. We will show, as an example, the case of $O_{ij}^p = 1$, i.e., Wigner-type $\delta$-function two-body interaction $(p=c)$. Then we have \begin{eqnarray} & &\sum_{i<j}^{N} \langle \phi_{\kappa} ({\bf s}_1,...,{\bf s}_N) \vert \delta({\bf r}_i-{\bf r}_j-{\bf r}) \vert \phi_{\kappa'}({{\bf s}'}_1,...,{{\bf s}'_N}) \rangle \nonumber \\ &=&\sum_{i<j}^{N} \sum_{k,l=1}^{N} \langle{\hat \varphi}_{{{\bf s}_i}{\sigma_i}{\tau_i}}^{\nu} {\hat\varphi}_{{{\bf s}_j}{\sigma_j}{\tau_j}}^{\nu} \vert \delta({\bf r}_i-{\bf r}_j-{\bf r}) \vert {\hat\varphi}_{{{{\bf s}'}_k}{\sigma'_k}{\tau'_k}}^{\nu} {\hat\varphi}_{{{{\bf s}'}_l}{\sigma'_l}{\tau'_l}}^{\nu} \rangle (-1)^{i+j+k+l} {\det} \lbrace B^{ijkl}\rbrace, \end{eqnarray} where $B^{ijkl}$ is obtained by omitting the $i$th and $j$th rows and the $k$th and $l$th columns of the matrix $B$. By substituting the explicit formula of the ingredients, we arrive at the same form as given in Eq. (25): \begin{eqnarray} & &\sum_{i<j}^{N} \langle \phi_{\kappa} ({\bf s}_1,...,{\bf s}_N) \vert \delta({\bf r}_i-{\bf r}_j-{\bf r}) \vert \phi_{\kappa'}({{\bf s}'}_1,...,{{\bf s}'_N}) \rangle \nonumber \\ &=&\bigl(\frac{\nu}{\pi}\bigr)^{3/2}{\rm e}^{-\nu r^2} \sum_{P}^{N!}C_P\sum_{i<j}^{N}{\rm e}^{-{1\over 2} \tilde{\bf s}( A_{P} + B_{P}^{(ij)}) {\bf s}+ {\bf d}_{P}^{(ij)}\cdot {\bf r}}, \end{eqnarray} with \begin{equation} {\bf d}_{P}^{(ij)}=\nu({\bf s}_i-{\bf s}_j + {\bf s}_{N+p_i}-{\bf s}_{N+p_j}). \end{equation} The coefficients $C_P$ and the matrices $A_P$ are the same as those that appear in Eq. (81). The matrix $B_P^{(ij)}$ is defined by \begin{eqnarray} {(B_P^{(ij)})}_{kl}={(B_P^{(ij)})}_{lk}=\left\{ \begin{array}{ll} \frac{\nu}{2} &\ \ \ (kl)=(ii), (jj), (N+p_i,N+p_i), (N+p_j,N+p_j),\\ & \ \ \ \ \ \ \ \ \ \ \ \ (i,N+p_i), (j,N+p_j). \\ -\frac{\nu}{2} & \ \ \ (kl)=(ij), (i,N+p_j), (j,N+p_i), (N+p_i,N+p_j). \\ 0 & \ \ \ {\rm otherwise}. \\ \end{array} \right. \end{eqnarray} \section*{Appendix D. Evaluation of the overlap of the correlated Gaussians} In this appendix we derive the overlap of the basis functions, defined in Eqs. (2) and (6), to illustrate the calculation of the matrix elements through the operations prescribed in Eq. (57). For the sake of simplicity we work out the case of overlap only, but the matrix elements of the kinetic and the potential energy are not much more involved either and these matrix elements can be calculated by repeating the steps detailed here. \par\indent In Eq. (57) the matrix element of an operator ${\cal O}$ between the correlated Gaussians are derived from the matrix elements between the Gaussian packets by applying suitably chosen operations. In the case of ${\cal O}=1$, upon substituting the matrix elements in the generator coordinate space (Eq. (41)), Eq. (57) reads as \begin{eqnarray} & &\langle {\cal A} \{f_{KLM}({\bf u},{\bf x},A){\chi}_{SM_S}{\cal X}_{TM_T}\} \vert {\cal A'} \{f_{K'L'M'}({\bf u}',{\bf x},A') {\chi}_{S'M'_S}{\cal X}_{T'M'_T}\}\rangle \nonumber \\ &\sim & \int\!\int d{\hat {\bf t}}d{\hat {\bf t}}' Y_{LM}({\hat {\bf t}})^{*} Y_{L'M'}({\hat {\bf t}}') \\ &\times& \left( {d^{\kappa+\kappa'} \over d\alpha^{\kappa}d\alpha'^{\kappa'}} {\rm e}^{-{1\over 2} \tilde{\bf T} C {\bf T}} \,\int d{{\bf S}} {\rm e}^{-{1\over 2} \tilde{{\bf S}} Q {{\bf S}}+\tilde{\bf T}{{\bf S}}} \sum_{i=1}^{n_o}C_i^{(o)} {\rm e}^{-{1\over 2}\tilde{\bf S} B_i^{(o)}{\bf S}} \right)_{\alpha=\alpha'=0 \atop t=t'=1}, \nonumber \end{eqnarray} with \begin{equation} \kappa=2K+L, \ \ \ \ \kappa'=2K'+L', \end{equation} where a constant factor is omitted for a moment. By integrating over the generator coordinates ${\bf S}$ with the use of Eq. (52), one readily obtains \begin{eqnarray} & &\langle {\cal A} \{f_{KLM}({\bf u},{\bf x},A){\chi}_{SM_S}{\cal X}_{TM_T}\} \vert {\cal A'} \{f_{K'L'M'}({\bf u}',{\bf x},A') {\chi}_{S'M'_S}{\cal X}_{T'M'_T}\}\rangle \nonumber \\ &\sim & \int\!\int d{\hat {\bf t}}d{\hat {\bf t}}' Y_{LM}({\hat {\bf t}})^{*} Y_{L'M'}({\hat {\bf t}}') \\ &\times& \left( {d^{\kappa+\kappa'} \over d\alpha^{\kappa}d\alpha'^{\kappa'}} \,\sum_{i=1}^{n_o}C_i^{(o)} \left({(2\pi)^{(2N-2)} \over {\rm det}(B_i^{(o)}+Q)}\right)^{3/2} {\rm e}^{p_i\alpha^2+{{p_i}'}\alpha'^2+q_i\alpha\alpha' {\bf t}\cdot{\bf t'}} \right)_{\alpha=\alpha'=0 \atop t=t'=1},\nonumber \end{eqnarray} where, by using Eqs. (54) and (55), we introduced the abbreviations \begin{eqnarray} & &p_i={1\over 2}\sum_{j,k=1}^{N-1} w_j ((B_i^{(o)}+Q)^{-1}-C)_{jk} w_k, \ \ \ {p_i}'={1\over 2}\sum_{j,k=N}^{2N-2} {w_j}' ((B_i^{(o)}+Q)^{-1}-C)_{jk} {w_k}' ,\nonumber \\ & &q_i=\sum_{j=1}^{N-1} \sum_{k=N}^{2N-2} w_j ((B_i^{(o)}+Q)^{-1}-C)_{jk} {w_k}' , \end{eqnarray} with \begin{equation} w_k=\sum_{j=1}^{N-1}(\Gamma C)^{-1}_{kj}u_j, \ \ \ \ {w_k}'=\sum_{j=N}^{2N-2}(\Gamma C')^{-1}_{kj}{u_j}' , \end{equation} in order to emphasize the $\alpha$- and $\alpha'$-dependence of the resulting expression. The differentiation with respect to $\alpha$ and $\alpha'$ can be promptly given by expanding the exponential functions into power series: \begin{eqnarray} & &{d^{\kappa+\kappa'} \over d\alpha^{\kappa}d{\alpha'}^{\kappa'}} {\rm e}^{p_i\alpha^2+{{p_i}'}\alpha'^2+q_i\alpha\alpha'{\bf t}\cdot{\bf t'}} \nonumber \\ & = & \sum_{j,j',k}^{\infty} { p_i^j{{p_i}'}^{j'}({q_i}{\bf t}\cdot{{\bf t}'})^k \over j!{j'}!k!} {(2j+k)!\over (2j+k-\kappa)!} {(2j'+k)!\over (2j'+k-\kappa')!} \alpha^{2j+k-\kappa} {\alpha'}^{2j'+k-\kappa'}, \end{eqnarray} and therefore, after putting $\alpha=\alpha'=0$, the calculation of Eq. (94) can be continued as \begin{eqnarray} & &\langle {\cal A} \{f_{KLM}({\bf u},{\bf x},A){\chi}_{SM_S}{\cal X}_{TM_T}\} \vert {\cal A'} \{f_{K'L'M'}({\bf u}',{\bf x},A') {\chi}_{S'M'_S}{\cal X}_{T'M'_T}\}\rangle \nonumber \\ &\sim & \int\!\int d{\hat {\bf t}}d{\hat {\bf t}}' Y_{LM}({\hat {\bf t}})^{*} Y_{L'M'}({\hat {\bf t}}') \\ &\times& \left(\sum_{i=1}^{n_o}C_i^{(o)} \left({(2\pi)^{2N-2} \over {\rm det}(B_i^{(o)}+Q)}\right)^{3/2} \kappa!\kappa'! \sum_{k} {{p_i^{\kappa-k\over 2}{{p_i}'}^{\kappa'-k\over 2}}{q_i}^k \over ({\kappa-k \over 2})! ({\kappa'-k\over 2})! k!} \left({\bf t}\cdot{\bf t}'\right)^{k}_{t=t'=1}\right), \nonumber \end{eqnarray} where the summation over $k$ runs from $0$ to max$(\kappa,\kappa')$ for those values that fulfill $(-1)^{k+\kappa}=1$ and $(-1)^{k+\kappa'}=1$. \par\indent The last step, the integration over the angles of ${\bf t}$ and ${\bf t}'$, can be accomplished by applying Eq. (49) to express the scalar product ${\bf t}\cdot {\bf t}'$, and then we get the final expression: \begin{eqnarray} & &\langle {\cal A} \{f_{KLM}({\bf u},{\bf x},A){\chi}_{SM_S}{\cal X}_{TM_T}\} \vert {\cal A'} \{f_{K'L'M'}({\bf u}',{\bf x},A') {\chi}_{S'M'_S}{\cal X}_{T'M'_T}\}\rangle \nonumber \\ &=&\delta_{LL'}\delta_{MM'} \frac{1}{B_{KL}B'_{K'L'}}\left(\frac {({\rm det}\Gamma)^3} {(4\pi)^{(N-1)}{\rm det}( \Gamma-A){\rm det}(\Gamma-A')} \right)^{3/2} \\ &\times& \left(\sum_{i=1}^{n_o}C_i^{(o)} \left({(2\pi)^{2N-2} \over {\rm det}(B_i^{(o)}+Q)}\right)^{3/2} \kappa!\kappa'! \sum_{k} {{p_i^{\kappa-k\over 2}{{p_i}'}^{\kappa'-k\over 2}}{q_i}^k \over ({\kappa-k \over 2})! ({\kappa'-k\over 2})! k!} B_{kL} \right), \nonumber \end{eqnarray} where the constant term previously suppressed is also included. The orthogonality of the matrix element in the spin and isospin quantum numbers is implicitly contained in the coefficients $C_i^{(o)}$. \par\indent The calculation of the matrix elements in the way described above is very simple. This fact becomes especially evident if one compares the work described above to the formidable task of the calculation of the matrix elements in the case where the function $\theta_{LM_L}({\bf x})$ is decomposed into partial waves of the relative coordinates as in Eq. (5). In fact, in that latter case one has to integrate over the angles of the relative coordinates and one has to cope with complicated angular momentum algebra. We note, however, that the calculation of the matrix element of the latter type poses no problem if the function $\theta_{LM_L}({\bf x})$ of Eq. (5) is expressed as a linear combination of the terms of Eq. (6) with appropriate ${\bf u}$-vectors. \par\indent All the matrix elements can be given in a similar closed analytic form and the numerical evaluation of the matrix elements as a function of the nonlinear parameters is therefore straightforward. The values $\kappa=2K\!\!+\!\!L$ and $\kappa'=2K'\!\!+\!\!L'$ are usually small in practical cases and the summation over $k$ is limited to just a few terms in Eq. (99). \bigskip \par\indent \par\indent This work was supported by OTKA grant No. T17298 (Hungary) and by a Grant-in-Aid for Scientific Research (No. 05243102 and No. 06640381) of the Ministry of Education, Science and Culture (Japan). K. V. gratefully acknowledges the hospitality of the RIKEN LINAC Laboratory and the support of the Science and Technology Agency of Japan, 1994. The authors are grateful for the use of RIKEN's VPP500 computer which made possible most of the calculations. Thanks are also due to the support of both Japan Society for the Promotion of Science and Hungarian Academy of Sciences, 1994-1995. The authors thank Prof. R. G. Lovas for careful reading of the manuscript and useful suggestions.
\section{Introduction} The analysis of the octet baryon masses in the framework of chiral perturbation theory already has a long history, see e.g. \cite{lang,pagels,juerg,gl82,liz,dobo,bkmz,lelu,samir,mary,dent,buergi}. In this paper, we present the results of a first calculation including {\it all} terms of order ${\cal O}(m_q^2)$, where $m_q$ is a generic symbol for any one of the light quark masses $m_{u,d,s}$. We work in the isospin limit $m_u = m_d$ and neglect the electromagnetic corrections. Previous investigations only considered mostly the so--called computable corrections of order $m_q^2$ or included some of the finite terms at this order. This, however, contradicts the spirit of chiral perturbation theory (CHPT) in that all terms at a given order have to be retained, see e.g. \cite{wein79,gl84,gl85}. In general, the quark mass expansion of the octet baryon masses takes the form \begin{equation} m = \, \, \krig{m} + \sum_q \, B_q \, m_q + \sum_q \, C_q \, m_q^{3/2} + \sum_q \, D_q \, m_q^2 + \ldots \label{massform} \end{equation} modulo logs. Here, $\krig{m}$ is the mass in the chiral limit of vanishing quark masses and the coefficients $B_q, C_q, D_q$ are state--dependent. Furthermore, they include contributions proprotional to some low--energy constants which appear beyond leading order in the effective Lagrangian. In this letter, we evaluate these coefficients for the octet baryons $N, \Lambda, \Sigma$ and $\Xi$. In addition, we also calculate the pion--nucleon $\sigma$--term which is intimately related to the quark mass expansion of the baryon masses \cite{juerg,bkmz,jms}. For some comprehensive reviews, see e.g. \cite{ulfrev,ecker,bkmrev,toni}. \section{Effective Lagrangian} To perform the calculations, we make use of the effective meson--baryon Lagrangian. Our notation is identical to the one used in \cite{bkmz} and we discuss here only the new terms. Denoting by $\phi$ the pseudoscalar Goldstone fields ($\pi, K, \eta$) and by $B$ the baryon octet, the effective Lagrangian takes the form \begin{equation} {\cal L}_{\rm eff} = {\cal L}_{\phi B}^{(1)} + {\cal L}_{\phi B}^{(2)} + {\cal L}_{\phi B}^{(4)} + {\cal L}_{\phi}^{(2)}+ {\cal L}_{\phi}^{(4)} \label{leff} \end{equation} where the chiral dimension $(i)$ counts the number of derivatives and/or meson mass insertions. The baryons are treated in the extreme non--relativistic limit \cite{jm,bkkm}, i.e. they are characterized by a four--velocity $v_\mu$. In this approach, there is a one--to--one correspondence between the expansion in small momenta and quark masses and the expansion in Goldstone boson loops, i.e. a consistent power counting scheme emerges. The form of the lowest order meson--baryon Lagrangian is standard, see e.g. \cite{bkmz}, and the meson Lagrangian is given in \cite{gl85}. The dimension two meson--baryon Lagrangian can be written as (we only enumerate the terms which contribute) \begin{equation} {\cal L}_{\phi B}^{(2)} = {\cal L}_{\phi B}^{(2, {\rm stand})} + \sum_{i=1}^{10} \, b_i \, O_i^{(2)} \, \, , \label{leff2} \end{equation} with the $O_i^{(2)}$ monomials in the fields of chiral dimension two. Typical forms are ${\rm Tr}(\bar{B} [ u_\mu , [ u^\mu,B]])$, ${\rm Tr}(\bar{B} [ v \cdot u , [ v \cdot u,B]])$ or $\bar{B} B {\rm Tr} (u_\mu u^\mu ) $, with $u_\mu = i u^\dagger \partial_\mu U u^\dagger$, $u = \sqrt{U}$ and $U=\exp(i \phi / F_P)$ collects the pseudoscalars. The form of ${\cal L}_{\phi B}^{(2, {\rm stand})}$ is \cite{bkmz}, \begin{equation} {\cal L}_{\phi B}^{(2, {\rm stand})} = b_D \, {\rm Tr}(\bar B \lbrace \chi_+ , B \rbrace ) + b_F \, {\rm Tr}(\bar B [\chi_+,B]) + b_0 \, {\rm Tr}(\bar B B) \, {\rm Tr}(\chi_+ ) \, \, , \label{leff2st} \end{equation} i.e. it contains three low--energy constants and $\chi_+ = u^\dagger \chi u^\dagger + u \chi^\dagger u$ is proportional to the quark mass matrix ${\cal M} ={\rm diag}(m_u,m_d,m_s)$ since $\chi = 2 B {\cal M}$. Here, $B = - \langle 0 | \bar{q} q | 0 \rangle / F_P^2$ and $F_P$ is the pseudoscalar decay constant. All low--energy constants in ${\cal L}_{\phi B}^{(2)}$ are finite. A subset of the $b_i$ has been estimated in \cite{norb} by analyzing kaon--nucleon scattering data. The splitting of the dimension two meson--baryon Lagrangian in Eq.(\ref{leff2}) is motivated by the fact that while the first three terms appear in tree and loop graphs, the latter ten only come in via loops. There are seven terms contributing at dimension four, \begin{equation} {\cal L}_{\phi B}^{(4)} = \sum_{i=1}^{7} \, d_i \, O_i^{(4)} \, \, , \label{leff4} \end{equation} with typical forms of the $O_i^{(4)}$ are $\bar{B}B {\rm Tr} (\chi_+^2)$ or ${\rm Tr}(\bar{B}[\chi_+ , [\chi_+ , B]] )$. At this stage, we take $m_u = m_d \ne m_s$. For $m_u \ne m_d$, there is an additional term in ${\cal L}_{\phi B}^{(4)}$. The explicit expressions for the various terms in Eqs.(\ref{leff2},\ref{leff4}) can be found in \cite{bora}. We point out that there are 20 a priori unkown constants. In addition, there are the $F$ and $D$ coupling constants (subject to the constraint $F+D = g_A =1.25$) from the lowest order Lagrangian ${\cal L}_{\phi B}^{(1)}$. What we have to calculate are all one--loop graphs with insertions from ${\cal L}_{\phi B}^{(1,2)}$ and tree graphs from ${\cal L}_{\phi B}^{(2,4)}$. We stress that we do not include the spin--3/2 decuplet in the effective field theory \cite{dobo}, but rather use these fields to estimate the pertinent low--energy constants (resonance saturation principle). We therefore strictly count in small quark masses and external momenta with no recourse to large $N_c$ arguments. \section{Baryon masses and pion--nucleon $\sigma$--term} The form of the terms $\sim m_q $ and $\sim m_q^{3/2}$ for the baryon masses and $\sigma_{\pi N} (0)$ is standard, we use here the same notation as Ref.\cite{bkmz}. The $q^4$ contribution to any octet baryon mass $m_B$ takes the form \begin{eqnarray} m_B^{(4)} & = & \epsilon_{1,B}^P \, M_P^4 + \epsilon_{2,B}^{PQ} \, M_P^2 \, M_Q^2 \nonumber \\ \quad & \, + & \epsilon_{3,B}^P \, M_P^4 \, \ln ( \frac{M_P^2}{\lambda^2}) + \epsilon_{4,B}^{PQ} \, M_P^2 \, M_Q^2 \, \ln ( \frac{ M_P^2}{\lambda^2} ) \, \, , \label{mB4} \end{eqnarray} with $P,Q = \pi , K, \eta$ and $\lambda$ the scale of dimensional regularization. The explicit form of the state--dependent prefactors $\epsilon_{i,B}$ can be found in Ref.\cite{bora}. Notice the appearance of mixed terms $\sim M_P^2 \, M_Q^2$ which were not considered in most existing investigations. The fourth order contribution to the baryon masses contains divergences proportional to (using dimensional regularization) \begin{equation} L = \frac{\lambda^{d-4}}{16 \pi^2} \biggl\lbrace \frac{1}{d-4} - \frac{1}{2}[\ln (4 \pi) +1 - \gamma_E] \biggr\rbrace \, \, , \label{L} \end{equation} with $\gamma_E = 0.572215$. These are renormalized by an appropriate choice of the low--energy constants $d_i$, \begin{equation} d_i = d_i^r (\lambda) + \Gamma_i \, L \, \, . \label{divd} \end{equation} In fact, all seven $d_i$ are divergent. The appearance of these divergences is in marked contrast to the $q^3$ calculation which is completely finite (in the heavy fermion approach). In what follows, we set $\lambda =1$~GeV. Similarly, the fourth order contribution to the pion--nucleon $\sigma$--term can be written as \begin{equation} \sigma_{\pi N}^{(4)}(0) = M_\pi^2 \, \biggl[ \, \epsilon_{1,\sigma}^P \, M_P^2 + \epsilon_{2,\sigma}^P \, M_P^2 \, \ln(\frac{M_P^2}{\lambda^2}) + \epsilon_{3,\sigma}^{PQ} \, M_P^2 \, \ln(\frac{M_Q^2}{\lambda^2}) \, \biggr] \, \, , \label{sigma4} \end{equation} with the $\epsilon_{i,\sigma}$ given in \cite{bora}. Here, the renormalization is somewhat more tricky. It can most efficiently performed in a basis of a linearly independent subset of the operators $dO_i^{(4)}/ dm_q$, $q=(u,d,s)$, as detailed in Ref.\cite{bora}. A good check on the rather lengthy expressions for the nucleon mass and $\sigma_{\pi N} (0)$ is given by the Feynman--Hellmann theorem, $\hat{m}(\partial m_N / \partial \hat{m}) = \sigma_{\pi N} (0)$, with $\hat m$ the average light quark mass, $\hat m = (m_u + m_d)/2$. We remark here that in contrast to the order $q^3$ calculation, the shift to the Cheng--Dashen point, $\sigma_{\pi N} (2M_\pi^2) - \sigma_{\pi N} (0)$, is no longer finite, i.e. there appear $t$--dependent divergences. We therefore do not consider this $\sigma$--term shift in what follows. We will also not discuss in detail the two kaon--nucleon $\sigma$--terms, $\sigma_{KN}^{(1,2)} (t)$, for similar reasons in this letter. \section{Resonance saturation} Clearly, we are not able to fix all the low--energy constants appearing in ${\cal L}_{\phi B}^{(2,4)}$ from data, even if we would resort to large $N_c$ arguments. We will therefore use the principle of resonance saturation to estimate these constants. This works very accurately in the meson sector \cite{reso,reso1,reso2}. In the baryon case, one has to account for excitations of meson ($R$) and baryon ($N^*$) resonances. One writes down the effective Lagrangian with these resonances chirally coupled to the Goldstones and the baryon octet, calculates the pertinent Feynman diagrams to the process under consideration and, finally, lets the resonance masses go to infinity (with fixed ratios of coupling constants to masses). This generates higher order terms in the effective meson--baryon Lagrangian with coefficients expressed in terms of a few known resonance parameters. Symbolically, we can write \begin{equation} \tilde{{\cal L}}_{\rm eff} [\, U,B,R,N^* \, ] \to {\cal L}_{\rm eff} [\, U,B \, ] \, \, . \end{equation} Here, there are two relevant contributions. One comes from the excitation of the spin-3/2 decuplet states and the second from t--channel scalar and vector meson excitations, cf. Fig.\ref{fig1}. It is important to stress that for the resonance contribution to the baryon masses, one has to involve Goldstone boson loops. This is different from the normal situation like e.g. in form factors or scattering processes. \begin{figure}[t] \hskip 1.2in \epsfysize=1.5in \epsffile{ma1.ps} \caption{\label{fig1} Resonance saturation for the masses. In (a), a baryon resonance (from the decuplet) is excited, whereas (b) represents the t--channel meson (scalar or vector) excitation. Solid and dashed lines denote the octet baryons and Goldstones, respectively.} \end{figure} Consider first the decuplet contribution. We treat these field relativistically and only at the last stage let the mass become very large. The pertinent interaction Lagrangian between the spin--3/2 fields (denoted by $\Delta$), the baryon octet and the Goldstones reads (suppressing SU(3) indices) \begin{equation} {\cal L}_{\Delta B \phi} = \frac{{\cal C}}{\sqrt{2} F_P} \, \bar{\Delta}^\nu \, T \, \Theta_{\nu \mu} (Z) \, B \, \partial^\mu \phi + {\rm h.c.} \, \, , \label{lmbd} \end{equation} with $T$ the standard $\frac{1}{2} \to \frac{3}{2}$ isospin transition operator, $F_P = 100$~MeV the (average) pseudeoscalar decay constant and ${\cal C} =1.5$ determined from the decays $\Delta \to B \pi$. The Dirac matrix operator $\Theta_{\mu \nu} (Z)$ is given by \begin{equation} \Theta_{\mu \nu} (Z) = g_{\mu \nu} - \biggl(Z + \frac{1}{2} \biggr) \, \gamma_\mu \, \gamma_\nu \, \, \, \, . \label{theta} \end{equation} For the off--shell parameter $Z$, we use $Z = -0.3$ from the determination of the $\Delta$ contribution to the $\pi N$ scattering volume $a_{33}$ \cite{armin}. The tadpole graphs shown in Fig.~1b with an intermediate vector meson only contribute at order $q^5$. This is evident in the conventional vector field formulation. In the tensor field formulation, the vertex ${\rm Tr}(V_{\mu \nu}u^\mu u^\nu)$ seems to lead to a contribution at order $q^4$. However, in a tadpole graph one needs the index contraction $\mu =\nu$ and thus has no term at order $q^4$. Matters are different for the scalars. Denoting by $S$ and $S_1$ the scalar octet and singlet with $M_S \simeq M_{S_1} \simeq 1$~GeV, respectively, the lowest order coupling to the baryon octet reads \begin{equation} {\cal L}_{SB}^{(0)} = D_S \, {\rm Tr}(\bar B \lbrace S,B \rbrace ) + F_S \, {\rm Tr}(\bar B [ S,B ] ) + D_{S_1} \,S_1 \, {\rm Tr}(\bar B B ) \label{sb} \end{equation} where the coupling constants $D_S$, $F_S$ and $D_{S_1}$ are of the order one.\footnote{In what follows, we neglect the singlet field.} In fact, there are no empirical data to really pin them down. From the decay pattern of the low--lying baryons one can, however, estimate these numbers to be small. We will generously vary them between zero and one. For the couplings of the scalars to the Goldstones, we use the notation of \cite{reso} and the parameters determined therein. Putting pieces together, all low--energy constants are expressed via resonance parameters and the baryon masses take the form \begin{eqnarray} m_B & = & \krig{m} + m_B^{(3)} + \lambda_B \, ({\krig{m}})^{-1} + \beta \, D_B^\beta + \delta \, D_B^\delta + \epsilon \, D_B^\epsilon + D_B^S \quad , \nonumber \\ \beta & = & -\frac{1}{96 \pi^2} \, \frac{m_\Delta^4}{{\krig{m}}^3} \, , \quad \delta= -\frac{\beta}{4{\krig{m}}^2} \, , \quad \epsilon = \frac{2}{3m_\Delta} (2Z^2 + Z -1 ) \, \, \, , \label{massres} \end{eqnarray} with $ m_B^{(3)}$ the contribution of ${\cal O}(m_q^{3/2})$ and $m_\Delta = 1.455$~GeV is the average decuplet mass. Notice that it is convenient to lump the ${\cal O}(m_q)$ and the ${\cal O}(m_q^2)$ corrections togther as it was done in Eq.(\ref{massres}) (in the $D_B^\beta$ and $D_B^S$). We have kept explict the baryon mass in the chiral limit. Of course, in the fourth order terms it could be substituted by the corresponding physical values. At second order, however, we would get a state--dependent shift, see e.g. \cite{bkmzas}, and we thus prefer to work with $\krig m$. Alternative representations for the baryon masses are given in \cite{bora}. Let us briefly explain the origin of the various terms in Eq.(\ref{massres}). The $\lambda$ contributions are tadpoles with $1/m$ insertions from ${\cal L}_{\phi B}^{(2)}$. Similarly, the $\beta$ and $\epsilon$ terms stem from tadpole graphs with insertions proportional to the low--energy constants $b_{0,D,F}$ and $b_i$, respectively. Note, however, that in the resonance exchange approximation not all of the ten $b_i$ are contributing. The $\delta$ terms subsume the contributions from ${\cal L}_{\phi B}^{(4)}$, these are proportional to the low--energy constants $d_i$. Finally, the terms of the type $D_B^S$ are the scalar meson contributions to the mass. They amount to a constant, state--dependent shift. These consist of terms of the types $\sim M_P^2$, $\sim M_P^4 \, \ln M_P^4$ and $\sim M_P^2 M_Q^2 \, \ln M_P^4$, see \cite{bora}. To be specific, we give the coefficients $D^{\beta,\delta,\epsilon}$ and $\lambda$ for the nucleon, \begin{eqnarray} & \lambda_N & = \frac{1}{8 \pi^2 F^2_P} \biggl\lbrace -\frac{3}{32} (D+F)^2 M_\pi^4 \ln \frac{M_\pi^2}{\lambda^2} - \frac{1}{96}(D - 3F)^2 M_\eta^4 \ln \frac{M_\eta^2}{\lambda^2} \nonumber \\ & \, & - \frac{1}{16}(D^2 - 2DF +3F^2) M_K^4 \ln \frac{M_K^2}{\lambda^2} \biggr\rbrace \, \, , \nonumber \\ & D_N^\beta & = D_N^{\beta,(2)} + D_N^{\beta,(4)} = \frac{{\cal C}^2 }{16 \pi^2 F^4_P} \biggl\lbrace (-4\pi^2 F_P^2)(M_K^2 + 4M_\pi^2) \nonumber \\ & + & \frac{27}{16} M_\pi^4 \ln \frac{M_\pi^2}{\lambda^2} +\frac{33}{24} M_K^4 \ln \frac{M_K^2}{\lambda^2} + \frac{7}{12} M_\eta^4 \ln \frac{M_\eta^2}{\lambda^2} \nonumber \\ & + & \biggl[ -\frac{11}{18}M_K^2 + \frac{43}{144}M_\pi^2 +\frac{1}{8}(D - 3F)^2 \bigl( \frac{M_K^2}{4}+M_\pi^2 \bigr) \biggr] M_\eta^2 \ln \frac{M_\eta^2}{\lambda^2} \nonumber \\ & + & \biggl[ \frac{9}{8}(D + F)^2 \bigl( \frac{M_K^2}{4} +M_\pi^2 \bigr) \biggr] M_\pi^2 \ln \frac{M_\pi^2}{\lambda^2} \nonumber \\ & + & \biggl[ -\frac{3}{4}(D - F)^2 M_K^2 + \frac{1}{96} (223D^2 - 318DF +351F^2) M_\pi^2 \biggr] M_K^2 \ln \frac{M_K^2}{\lambda^2} \biggr\rbrace \, \, , \nonumber \\ & D_N^\delta & = \frac{{\cal C}^2}{2F_P^2} \biggl\lbrace -\frac{415}{288} M_\pi^4 + \frac{83}{72} M_\pi^2 M_K^2 - \frac{31}{72} M_K^4 \biggr\rbrace \, \, , \nonumber \\ & D_N^\epsilon & = \frac{{\cal C}^2}{8 \pi^2 F_P^2} \biggl\lbrace - \frac{1}{2} M_\pi^4 \ln \frac{M_\pi^2}{\lambda^2} - \frac{1}{8} M_K^4 \ln \frac{M_K^2}{\lambda^2} \biggr\rbrace \, \, . \label{DN} \end{eqnarray} The corresponding coefficients for the $\Lambda$, $\Sigma$ and $\Xi$ and also the $D_B^S$ can be found in \cite{bora}. The tree contribution from ${\cal L}_{\phi B}^{(2, {\rm stand})}$ is subsumed in the $D_B^\beta$. The numerical values of the $\lambda_B$, $D^\beta_B, D^\delta_B$, $D^\epsilon_B$ and $D_B^S$ are given in table~1 (using $F = 0.5$, $D=0.75$ and $c_{d,m}$ from \cite{reso}). \begin{table}[h] \begin{tabular}{|c|ccccc|} \hline B & $\lambda_B$ [GeV$^{2}$] & $D^\beta_B$ & $D^\delta_B$ [GeV$^{2}$] & $D^\epsilon_B$ [GeV$^{2}$] & $D_B^S$ [GeV] \\ \hline $N$ & 0.0049 & $-$39.875 & $-$2.3617 & 0.0321 & $-0.017 \, D_S + 0.061 \, F_S$ \\ $\Lambda$ & 0.0179 & $-$68.044 & 1.0831 & 0.0617 & $-0.047 \, D_S - 0.010 \, F_S$ \\ $\Sigma $ & 0.0144 & $-$132.90 & $-$11.273 & 0.1422 & $+0.051 \, D_S + 0.010 \, F_S$ \\ $\Xi$ & 0.0216 & $-$126.46 & $-$5.4799 & 0.1324 & $-0.037 \, D_S - 0.082 \, F_S$ \\ \hline \end{tabular} \medskip \caption{Numerical values of the state--dependent coefficients in Eq.(\protect\ref{massres}). The $D_B^\beta$ are dimensionless. The $D_B^S$ are for $c_d$, $c_m$ from \protect\cite{reso} and $M_S =1$ GeV.} \end{table} We see that the dominant terms at ${\cal O}(m_q^2)$ are indeed the tadpole graphs with an insertion from ${\cal L}_{\phi B}^{(2, {\rm stand})}$ (this holds for the masses but not for $\sigma_{\pi N}(0)$). It is also instructive to compare the values we find from resonance exchange with the ones previously determined from KN scattering data. We have transformed the results of Ref.\cite{norb} into our notation. As can be seen from table~2, most (but not all) coefficients agree in sign and magnitude. \begin{table}[bht] \begin{tabular}{|c|ccccccc|} \hline & $b_1$ & $b_2$ & $b_3$ & $b_8$ & $b_0$ & $b_F$ & $b_D$ \\ \hline Reso. $(D_S=F_S=0)$ & 0.084 & $-$0.144 & 0.108 & $-$0.216 & $-$0.738 & $-$0.264 & 0.317 \\ Reso. $(D_S=F_S=1)$ & 0.100 & $-$0.111 & 0.125 & $-$0.239 & $-$0.733 & $-$0.208 & 0.345 \\ Ref.\cite{norb} & 0.044 & $-$0.145 & $-$0.054 & $-$0.165 & $-$0.493 & $-$0.213 & 0.066 \\ \hline \end{tabular} \medskip \caption{Some low--energy constants from ${\cal L}_{\phi B}^{(2)}$ in GeV$^{-1}$. In the first row, only the decuplet contribution is given. In the second row, scalar meson exchange is added. \end{table} Note, however, that this is only a subset of the coefficients considered in this work. The full list will be given in \cite{bora}. We remark that the procedure used in \cite{norb} involves the summation of arbitrary high orders via a Lippmann--Schwinger equation and is thus afflicted with some uncertainty not controled within CHPT. \section{Results and discussion} The only free parameter in the formula for the baryon masses, Eq.(\ref{massres}), is the mass of the baryons in the chiral limit, $\krig m$, since all low--energy constants are fixed in terms of resonance parameters. In particular, in contrast to Ref.\cite{bkmz}, the parameters $b_0$, $b_D$ and $b_F$ are no longer free. Also, at quadratic order in the quark masses the ambiguity between $\krig m$ and $b_0$ is resolved, it is not necessary to involve any one of the $\sigma$--terms in the fitting procedure \cite{liz,bkmz}. In fact, one can not find one single value of $\krig m$ to fit all four octet masses, $m_N = 0.9389$~GeV, $m_\Lambda = 1.1156$~GeV, $m_\Sigma = 1.1931$~GeV and $m_\Xi = 1.3181$~GeV exactly. We therefore fit these masses individually and average the corresponding values for $\krig m$. The contribution from scalar meson exchange only enters the uncertainties of the numbers given. This is justified since the numerical values for the couplings $F_S$ and $D_S$ are supposedly small. A more thorough discussion on this point can be found in \cite{bora}. We have $\krig{m}_N = 711 \, \, {\rm MeV}$, $\krig{m}_\Lambda = 679 \, \, {\rm MeV}$, $\krig{m}_\Sigma = 877 \, \, {\rm MeV}$, $\krig{m}_\Xi = 728 \, \, {\rm MeV}$, with the following average \begin{equation} \krig{m} = 749 \pm 125 \, \, {\rm MeV}\,\, . \quad \label{mkrig} \end{equation} This number is compatible with the one found in the analysis of the pion--nucleon $\sigma$--term, where approximately 130 MeV to the nucleon mass were attributed to the strange matrix element $m_s \langle p | \bar s s|p \rangle$ (with a sizeable uncertainty) \cite{gls}. The spread of the various values is a good measure of the uncertainties related to this complete $q^4$ calculation. Let us take a closer look at the quark mass expansion of the nucleon mass, in the notation of Eq.(\ref{massform}), \begin{equation} m_N = 711 + 202 - 272 + 298 \, \, \, {\rm MeV} = 939 \, \, {\rm MeV} \, \, . \label{mnuc} \end{equation} This looks similar for the other octet baryons. We conclude that the quark mass corrections of order $m_q$, $m_q^{3/2}$ and $m_q^2$ are all of the same size.\footnote{Note that for certain values of the scalar couplings, the fourth order term $m_N^{(4)}$ can be significantly smaller. However, the nucleon mass is much more sensitive to the scalar contribution than the other octet masses. } In Ref.\cite{jms}, it was argued that only the leading non--analytic corrections (LNAC) $\sim m_q^{3/2}$ are large and that further terms like the ones $\sim m_q^2$ are modestly small, of the order of 100 MeV. This would amount to an expansion in $M_K^2/(4 \pi F_P)^2 \sim 1/6$ with a large leading term. This expectation is not borne out by our results, the next corrections are as large as the LNACs. These findings agree with the meson cloud model calculation of Gasser \cite{juerg}. A last remark about the baryon masses concerns the deviation from the Gell-Mann-Okubo relation, $\Delta_{\rm GMO} = (3m_\Lambda+m_\Sigma-2m_N-2m_\Xi)/4$ which empirically is about 6.5 MeV. We find $\Delta_{\rm GMO} = 31$ MeV, which is larger in magnitude than the value found in \cite{liz,dobo}. We remark that in our case the decuplet contribution is contained in the $m_q^2$ contributions and not in the $m_q^{3/2}$ as in \cite{liz}. Therefore, in our case, $\Delta_{\rm GMO}$ is dominated by the $m_q^2$ piece. The sizeable uncertainty in the chiral limit masses, Eq.(\ref{mkrig}), does not allow for a very accurate statement about this very small quantity. It is also very sensitive to the scalar couplings. To get a better handle on this issue, one either has to be able to fix all pertinent low--energy constants at order $m_q^2$ from data or improve upon the resonance saturation estimate used here by including e.g. the mass splitting within the decuplet and the SU(3) breaking of the decuplet-octet-meson couplings. A better understanding of this topic is, of course, at the heart of the determination of the quark mass ratio $R = (m_d-m_u)/(m_s - \hat m )$ from the baryon masses (once the electromagnetic corrections have been included). The pion--nucleon $\sigma$--term is completely fixed. Using for $\krig m$ its average, we find (no scalar resonance contribution) \begin{equation} \sigma_{\pi N} (0) = 48 \, \, {\rm MeV} \, \, , \label{sigval} \end{equation} with an uncertainty of about $\pm 10$ MeV due to the spread in $\krig m$. This number compares favourably with the one found in Ref.\cite{gls}, $\sigma_{\pi N} (0) = 44 \pm 9$~MeV. We stress that this result reflects a very non--trivial consistency for the complete calculation to quadratic order in the quark masses using the resonance saturation principle in the scalar sector. The additional contribution from the scalar meson exchange is accounted for in the $\pm 10$ MeV uncertainty. To be specific, we have $\sigma_{\pi N} (0) = (48 + D_S \cdot 0.5 - F_S \cdot 2.0) \, \, {\rm MeV}$. It is furthermore instructive to disentangle the various contributions to $\sigma_{\pi N} (0)$ of order $q^2$, $q^3$ and $q^4$, respectively, \begin{equation} \sigma_{\pi N} (0) = 54 - 33 + 27 \, \, \, {\rm MeV} = 48 \, \, {\rm MeV} \, \, , \label{signo} \end{equation} which shows a moderate convergence, i.e. the terms of increasing order become successively smaller. Still, the $q^4$ contribution is important. Also, at this order no $\pi \pi$ rescattering effects are included. We notice that using the values for $b_{0,D,F}$ and $b_i$ as determined in Ref.\cite{norb} leads to a much increased fourth order contribution. \section{Summary and outlook} In this paper, we have used heavy baryon chiral perturbation theory to calculate the octet baryon masses to quadratic order in the quark masses, including 20 local operators with unknown coeffcients. These low--energy constants were fixed by resonance exchange. The dominant contribution comes indeed from the excitation of the spin--3/2 decuplet fields. Tadpole graphs with scalar meson exchange only lead to small corrections. This left us with one free parameter, the baryon mass in the chiral limit, which could be determined within 18\% accuracy, and is compatible with the value inferred for the strange matrix--element $m_s \langle p | \bar s s | p \rangle$ in Ref.\cite{gls}. Furthermore, the pion--nucleon $\sigma$--term comes out surprisingly close to its empirical value, $\sigma_{\pi N} (0) = 48$~MeV, with an uncertainty of about $\pm 10$ MeV. This first exploratory $q^4$ study of the three flavor scalar sector of baryon CHPT points towards a significant improvement compared to previous investigations which were mostly confined to so--called ``computable'' corrections and/or fitted a few of the pertinent low--energy constants. However, the calculation is not yet accurate enough to determine the quark mass ratio $R$ reliably from the octet masses. Furthermore, we did not address the kaon--nucleon $\sigma$--terms, the corresponding shifts to the pertinent Cheng-Dashen points together with the strangeness content of the proton here. We will come back to these topics in Ref.\cite{bora}. \section*{Acknowledgements} We are grateful to Daniel Wyler for a very useful remark.
\section{Introduction} Self-avoiding random walk models appeared in chemical physics as models for long polymer chains. Roughly speaking, a polymer is composed of a large number of monomers which are linked together randomly but cannot overlap. This feature is modelled by a self-repulsion term. Let ${\bf A}$ be a d-dimensional hypercubic lattice, typically the integer lattice ${\bf Z}^d$ (or a finite subset of ${\bf Z}^d$). Its elements are called sites, oriented pairs of sites are called steps. A walk $w$ on ${\bf A}$ is an ordered sequence $(w(0),w(1),...,w(k))$ of sites in ${\bf A}$; $k\geq 0$, $k$ being the length of the walk \cite{Fer}. Thus, a simple random walk on ${\bf A}$, starting at $w(0)\in {\bf Z}^d$, is a stochastic process indexed by the non-negative integers.\ To state a self-repulsion term, Flory \cite{Fr} used the self-avoiding walk (SAW). A self-avoiding walk of length $n$ is a simple random walk which visits no site more than once. Although this simple model possesses some qualitative features of polymers, turns out to be very difficult for obtaining rigorous results. Instead of using SAW we take advantage of the several measures on random walks which favor self-avoiding walks.\ In this paper we study weakly SARW (so called Domb-Joyce model or self-repellent walk). This is a measure on the set of simple walks in which self-intersections are discouraged but not forbidden. Here, double intersections of random walks are penalized by a factor $e^{-\lambda \mbox{ (measure of (self-)intersection)}}$, $\lambda>0$ being a small constant. This factor is needed to make the process tend to avoid itself. Thus, probabilities of the random walks are modified by a measure of (self-)intersection inside the lattice. This measure is written in terms of waiting (or local) times for the process ($t_i$). Here, as was done in reference [5], the measure is $\sum_{i<j}t_it_j$, for all local times in the\newline (self-)intersections of the walks on the lattice, which can be either hypercubic or hierarchical. If in this model the lattice is the hypercubic lattice, then the state space is ${\bf Z}^d$. To understand this model, we use a hierarchical lattice which state space is defined in next section. We want to stress that the only feature of the model that depends upon the lattice used (hypercubic or hierarchical) is the state space in this. Thus, definitions for the weakly SARW in section 4 (interaction energy in Theorem 4) are equivalent in both cases, but with different state space in the lattice. We develop our method on a hierarchical lattice because they have the feature that the renormalization-group map is particulary simple, which is not the case on the hypercubic lattice. We believe that the results of this procedure extend to weakly SARW on the hypercubic lattice.\ The hierarchical models introduced by Dyson \cite{Dy} feature a simple \newline renormalization-group transformation. This can easily be seen in related literature $\cite{Ga}$. We would like to understand the logarithmic correction that appears in the hypercubic lattice, d=4, for the end-to-end distance of the weakly SARW. The real space renormalization-group map we develop here, is factorizable only in terms of a hierarchical lattice. So, we present a method in which $A$ is labeled in terms of a hierarchical metric space, from this an easy realization of the map is followed. In the integer lattice ${\bf Z}^d$, this is not true and technical problems arise.\ We use a hierarchical lattice where the points are labeled by elements of a countable, abelian group ${\it G}$ with ultrametric $\delta$; i.e. the metric space $({\it G},\delta)$ is hierarchical. The hierarchical structure of this metric space induces a renormalization-group map that is ``local"; i.e. instead of studying the space of random functions on the whole lattice, we can descend to the study of random functions on L-blocks (cosets of {\it G}) \cite{Ev}. This simplifying feature was used by Brydges, Evans and Imbrie \cite{Ev} to prove (in the $\lambda \phi^4$ Grassmann valued field representation for a weakly SARW that penalizes the (self-)intersection of two random walks) that the introduction of a sufficiently weakly self-avoidance interaction does not change the decay of the Green's function for a particular L\'{e}vy process (continuous time random walk), \cite{Ev} when d=4, provided the mass is introduced critically.\ A rigorous proof of the end-to-end distance for the weakly SARW, $d=4$, on the hierarchical lattice, has recently been reported \cite{Im}. This was done in the field theoretical approach by means of controlling the interacting Green's function and inverting the Laplace transform.\ Low dimensional models are the most interesting from physical viewpoint, but rigorous results are difficult of being obtained. One major result is the proof that in high dimensions the exponents of weakly SARW take the ``mean-field" value \cite{Ma}. In this context ``lace expansion" is one of the most successful tools. Contrary to what has been done from this method, in this paper we develop a map on real space which is full of the physical intuition needed for being applied on some other cases not yet solved.\ Weakly SARW exhibits logarithmic corrections for physically meaningful magnitudes in the critical dimension of the model, i.e. d=4. We study the end-to-end distance for weakly SARW that penalizes the double crossing of walks in $d=4$. A probabilistic meaning is given to the exponent of this logarithmic correction. In this paper we present an heuristic space-time renormalization-group argument to show that the end-to-end distance of a weakly self-avoiding random walk (SARW) on the hierarchical lattice, that penalizes the (self-)intersection of two walks in $d=4$, is asymptotic to a constant times $T^{\frac{1}{2}}log^{\frac{1}{8}}T$ as T tends to infinity, T being the total time for the walk. This has already been conjectured before \cite{Br}. This is the testing ground to check that our map reproduces previously known results with physical improved intuition. The weakly self-avoiding random walk model that penalizes the (self-)intersection of two random walks is in the same universality class as the perfect self-avoiding random walk model; therefore, the same logarithmic correction for the end-to-end distance is expected to hold for both cases regardless the details of the state space used. Real space renormalization-group methods have proved to be useful in the study of a wide class of phenomena.\ Since our method is intended to provide an alternative way, full of physical intuition, to renormalize random walk models on a hierarchical lattice; we study weakly SARW in $d=4$ just as a testing ground. We consider our method suitable of being directly applied on kinetically growing measure models, discrete version. Kinetically growing measure models are produced from consistent measures and are the natural framework for random walks provided these are seen like stochastic processes. Myopic self-avoiding walk (a model for adsorption of linear polymers on surfaces \cite{Bu}), infinite self-avoiding walk and Laplacian (or loop-erased) random walk \cite{La} are examples of kinetically growing measure models. These models have been studied mainly in the field theoretical framework where has been shown \cite{De} that some contributions neglected in the derivation of the continuum problem from the discrete version of the model might be essential in determining the asymptotic behaviour of the model analytically. In fact, usually the corresponding action (in the continuum limit) is not fully renormalizable. We address this problem by presenting a real space renormalization-group suitable of been applied on discrete models. The metod can be used for obtaining both, rigorous and heuristic results for these models.\ We hope this paper could be interesting for various readerships. We intend to brydge a gap between probabilistic approaches and field theoretical ones, thereby providing a new probabilistic meaning to critical and asymptotic exponents. The method we develop in this paper can be used also as an intuitive mean to search new exact results, which then remain to be proven rigorously within the framework presented here, or by other means. The renormalization scheme constructed here is not considered or known in the literature.\ This paper is organized as follows; in Section 2 we present the hierarchical lattice and the L\'{e}vy process we study, on the space of simple random walks. In Section 3 we define the renormalization-group map on the hierarchical lattice and prove that two particular probabilities, functions of random walks, flow to fixed forms after applying the map. In Section 4 we apply the renormalization-group map to the weakly SARW model that penalizes the intersection of two random walks on the hierarchical lattice. In Section 5 we present an heuristic proof for the asymptotic behavior of the end-to-end distance for the weakly SARW on the hierarchical lattice, d=4. Although this is an heuristic proof, it helps in understanding the way the map can be used and gives a new probabilistic meaning to the exponent of the logarithmic correction. Sections 3, 4 and 5 involve important results of this paper, being Theorem 4 the main result reported in the paper and an important step to obtain heuristically the end-to-end distance for the weakly SARW. Even more, we claim this Theorem to be a random walk version of the field theoretical approach in reference [5] with improved physical intuition. Finally, we summarize.\ \section{The hierarchical random walk.} The hierarchical lattice used in this paper was recently introduced by Brydges, Evans and Imbrie \cite{Ev} \cite{Im}. Here we are presenting a slight variant of the model in reference [5].\ Fix an integer $L\geq 2$. Hereafter, the points of the lattice ${\bf A}$ are labeled by elements of the countable abelian group ${\it G}=\oplus ^{\infty}_{k=0}{\bf Z}_{L^d}$, d being the dimension of the lattice. Through the paper the abelian group ${\it G}=\oplus ^{\infty}_{k=0}{\bf Z}_{L^d}$ replaces ${\bf A}$.\ An element $X$ in ${\it G}$ is an infinite sequence $$ X\equiv (...,X_k,...,X_2, X_1, X_0 )\;\; ; \mbox{$X_i \in {\bf Z}_{L^d}$ thus $X\in {\it G}=\oplus^{\infty}_{k=0} {\bf Z}_{L^d}$}, $$ where only finitely many $X_i$ are non-zero.\ Let us define subgroups \begin{equation} \{0\}={\it G}_0\subset {\it G}_1\subset ...\subset {\it G} \;\;\;\mbox{where } {\it G}_k=\{X\in {\it G}| X_i=0, i\geq k \}, \end{equation} and the norm $|\cdot|$ as \begin{equation} |X|=\left\{\begin{array}{cc} 0 & \mbox{ if $X=0$} \\ L^p & \mbox{where $p=\inf\{k| X\in {\it G}_k\}$ if $X\neq 0$}. \end{array} \right. \end{equation} Then, the map $\delta:(X, Y)\rightarrow |X-Y|$ defines a metric on ${\it G}$. In this metric the subgroups ${\it G}_k$ are balls $|X|\leq L^k$ containing $L^{dk}$ points. Here the operation + (hence - as well) is defined componentwise.\ In Figure 1 we have described two examples for $L=2$, a one-dimensional hierarchical lattice (Figure 1.a)) already presented in reference [5] and a two-dimensional hierarchical lattice (Figure 1.b)). In these Figures we depict ${\it G}_k$ cosets, a way to calculate distances among points, and the concept of scales for each example.\ The metric defined by eq(2) satisfies a stronger condition than the triangle inequality, namely \begin{equation} |X+Y|\leq \mbox{ Max}(|X|,|Y|). \end{equation} This {\it non-archimedean} property implies that every triangle is isosceles and that every point interior to a ball can be considered its center. Moreover, balls of radius L are the same as balls of diameter L, and are the same as ${\it G}_1$ cosets. From inequality (3), it is clear that the metric introduced is an ultrametric and confers the hierarchical structure. Strictly speaking, it is only the metric space $({\it G},\delta)$ that is hierarchical. Here, ultrametric appears naturaly as a property of polynomials. It can be shown that ${\it G}_k$ represents polynomials of degree $k$ on a formal basis.\ Let us now introduce the L\'{e}vy process we propose in this paper. The elements of the lattice ${\it G}$ are called sites; unoriented pairs $\{X,Y\}$ of sites in ${\it G}$ with $X\neq Y$ are called bonds; oriented pairs $(X,Y)$ are called steps (or jumps) with initial site $X$ and final site $Y$. Let us define the L\'{e}vy process \cite{Ev}($\equiv $ continuous time random walk), $w$, as an ordered sequence of sites in {\it G}; \begin{equation} (w(t_0),...,w(t_0+...+t_n))\;\;,w(t_0+...+t_i)=X_i\in {\it G}, \;\;T=\sum ^n_{i=0}t_i ,\; n\geq 0 \end{equation} where $t_i$ is the time spent in $X_i\in {\it G}$ (waiting time at $X_i$) and T, fixed, is the running time for the process. For convenience we take $X_0=0$. The support of the walk $w$ is defined by \begin{equation} supp(w)=\{X\in {\it G} | w(t_0,...,t_j)=X \mbox{ for some j}\}, \end{equation} for any $w$. The random walk we are dealing with is not the nearest neighbor random walk on the lattice, provided we mean neighbourhood with respect to the ultrametric distance $\delta$ previously defined.\ If we compare this L\'evy process with the simple random walk in the hypercubic lattice defined in the Introduction, we see that in this Section we construct a different stochastic process. Here, time is continuous and the path of the walk is given on ${\it G}$ (in the hypercubic lattice, the walk was indexed by non-negative integers, instead of waiting times, with path on ${\bf Z}^d$). Besides, here we fix the running time for the process. This corresponds to walks of fixed length, feature that was not imposed on simple random walks on the hypercubic lattice.\ We suppose the L\'{e}vy process in the hierarchical lattice has a probability $rdt$ (r is the jumping rate) of making a step in time $(t, dt)$ and, given that it jumps, the probability of jumping from X to Y is $q(X,Y)$. Thus, the process, conditioned to $n$ jumps spaced by times $t_0,t_1,...,t_n$, has a probability density \begin{equation} P(w)=r^ne^{-rT}\prod^{n-1}_{i=0}q(X_{i+1},X_{i}) \;\;\mbox{, where $T=\sum^{n}_{i=0}t_i$.} \end{equation} Define $Dw$ by $$ \int(\cdot ) Dw= \sum_{n}\;\;\sum_{\left[ X_{i}\right]^{n}_{i=0}} \;\;\;\int^{T}_{t_i=0}\;\;\prod^{n}_{i=0}\;\;dt_{i} \delta(\sum^{n}_{i=0}t_{i}-T)(\cdot ). $$ {}From this and eq(6) it is straightforward to obtain $$ \int P(w)Dw= \left\langle \sum_{\left[ X_{i}\right]^{n-1}_{i=0}} \prod^{n-1}_{i=0}q(X_{i+1},X_{i}) \right\rangle_{Poisson}=1, $$ $\left\langle (\cdot) \right\rangle$ being the expectation of ($\cdot$).\ Here $\prod^{n-1}_{i=0}q(X_{i+1},X_{i})$ has been normalized on ${\it G}$ and we have used $$ \int^T_0\prod^n_{i=0}r^ne^{-rt_i}\prod^n_{j=0}dt_j\delta \left(\sum^n_{j=0}t_j-T\right)= $$ $$ r^ne^{-rT}\int^T_0\prod^n_{i=0}dt_i\delta \left(\sum^n_{i=0}t_i-T\right)=\frac{\left(rt^n\right)}{n!}e^{-rT}. $$ \section{The renormalization-group map.} To start with, let us introduce a renormalization-group map on the lattice; $R(X_i)=LX'_i$ where $X_i\in {\it G}$ and $LX'_{i}\in {\it G}'$ =${\it G}/{\it G}_1\sim {\it G}$; i.e. the renormalized lattice ${\it G}'$ is isomorphic to the original lattice ${\it G}$.\ {}From this renormalization-group map we construct $R(w)=w'$, from $w$ above as defined, to $w'$. Here, $w'$ is the following ordered sequence of sites in ${\it G}'= {\it G}/{\it G}_1\approx {\it G}$; \begin{equation} (w'(t'_0),...,w'(t'_0+...+t'_k))\;\;,\mbox{ where} \end{equation} $$ w'(t'_0,...,t'_{i'})=X'_{i'}\in {\it G},\;\;T'=\sum ^k_{i'=0}t'_i , \;0\leq k\leq n, \;\;T=L^{\beta}T'. $$ R maps $w(t_0),w(t_0+t_1),...,w(t_0+...+t_n)$ to cosets $w(t_0)+{\it G}_1,w(t_0+t_1)+{\it G}_1,...,w(t_0+...+t_n)+{\it G}_1$. If two or more successive cosets in the image are the same, they are listed only as one site in $w'(t'_0),...,w(t'_0+...+t'_k)$, and the times $t'_j$ are sums of the corresponding $t_j$ for which successive cosets are the same, rescaled by $L^{\beta}$. For $\beta=2$, we are dealing with normal diffusion, in case $\beta<2$ with superdiffusion, and subdiffusion for $\beta>2$. Additionally, the renormalization-group maps each ${\it G}_1$ coset to the center of the ball and rescales by $L$. In reference [5], $\beta$ is set to 2.\ The renormalization group map, applied to functions of the hierarchical random walk, preserves locality \cite{Ev}. Thus, if $F(w)=\prod_{X_{i}\in {\it G}}f(w(t)=X_{i})$, the effect of the renormalization-group map on $F(w)$ can be studied as the product, for all elements of the group ${\it G}/{\it G}_1={\it G}'\sim {\it G}$, of the images of the renormalization-group map of $f(w)$ in the ${\it G}_1$ coset \cite{Ev}. This can be seen as follows; $\prod_{X_{i}\in {\it G}}$ splits into two parts; the first part is $\prod_{X'_{i'}\in {\it G}/{\it G}_1}$ which corresponds to $\prod_{X'_{i'}\in {\it G}}$. The second part is $\prod_{X_{i}\in ({\it G})_{X'_{i'}}}$ and stands for the $L$-block (${\it G}_1$ coset) of the lattice which, under the renormalization-group transformation, maps to $LX'_{i'}$, $(0\leq i'\leq k)$. The geometrical interpretation of this is quite simple. The renormalization-group map applied on $F(w)$ in ${\it G}$ splits into that of $f(w(t)=X_{i})$ in the contracting ${\it G}_1$ cosets, multiplied by the whole ${\it G}/{\it G}_1$ group. To study the renormalization of $F(w)$ in ${\it G}$ we descend to analize the renormalization-group action on $f(w(t)=X_{i})$ in the contracting ${\it G}_1$ coset.\ In Figure 2 we present the lattice ${\it G}$. On this, walks with fixed topology in ${\it G}/{\it G_1}$ are depicted. In this example three different types of local topologies inside the ${\it G}_1$ cosets are illustrated. Once the renormalization-group map is applied, the renormalized lattice and random walk are also shown. The renormalized random walk $w'$ visits each type of ${\it G}_1$ cosets once, twice and three times respectively. In the renormalized lattice we show the contracted ${\it G}_1$ cosets. The particular fixed topology chosen for $w'$ is one in the class of the simplest cases studied by our method. Here is clear what is meant by factorization and locality of the renormalization-group map. The map factorizes into two terms. Roughly speaking, the first term corresponds to events inside $L$-blocks or ${\it G}_1$ cosets and the second term corresponds to events outside the $L$-block (${\it G}_1$ cosets); therefore in ${\it G}/{\it G}_1$=$ {\it G}'\sim {\it G}$. Moreover, for obtaining the flow of the interaction constant, the map descends to study events in ${\it G}_1$ cosets; thereby preserving locality.\ We can now work out probabilities at the $(p+1)^{th}$ stage in the renormalization provided only that we know the probabilities at the $p^{th}$ stage. We sum over the probabilities of all the walks $w^{(p)}$ consistent with a fixed walk $w^{(p+1)}$ in accordance with the following;\ {\bf Definition. Let $R(w)=w'$ be the renormalization-group map, above as stated. Then \begin{equation} P'(w')= L^{\beta k}\int Dw P(w)\chi (R(w)=w') \end{equation} for any probability $P(w)$ where the running time for the process $T=\sum^{n}_{i=0}t_{i}$ is fixed}.\ In this definition, R is a renormalization-group transformation that maps a density $P(w)$ to a new one, $P'(w')$, on rescaled coarse walks. Besides, $\chi(c)$ is the characteristic function of the condition c.\ Let $P(w)=\prod_{X\in {\it G}}p(w(t)=X)$ then\newline $P'(w')$=$L^{\beta k}\int Dw\prod_{X\in {\it G}}p(w(t)=X)\chi (R(w)=w')$ . {}From this and factorization properties in the hierarchical lattice follows $$ P'(w')=\prod_{X'\in{\it G}}\left\{ L^{\beta k}\int Dw\prod_{X\in ({\it G}_1)_{X'}}p(w(t)=X)\chi(R(w)=w')\right\} $$ $$ = \prod_{X'\in {\it G}}p'(w'(t')=X'), $$ that proves in this case the statement of preservation of locality as above given for $F(w)$. Eq(17) and eq(21) are examples for suitable $P(w)$.\ Eq(8) corresponds precisely to \begin{equation} P'(w')= L^{\beta k}\;\sum_{\left[ n_{i'}\right]^{k}_{i'=0}} \;\sum_{\left[ X_{i}\right]^{n}_{i=0}} \int \prod^{n}_{i=0}dt_{i}\; \prod^{k}_{j'=0}\; \delta (\sum^{m_{j'}}_{{i}=m_{j'-1}+1}\;\;t_{i}- L^{\beta }t'_{j'})\times \end{equation} $$ \times \prod^{k}_{j'=0}\; \prod^{m_{j'}}_{{i}=m_{j'-1}+1}\; \chi (X_{i}\in LX'_{j'}\;\;)P(w). $$ Hereafter \begin{equation} m_{j'}=\sum^{j'}_{i'=0}n_{i'}+j' \;\;\mbox{ and} \end{equation} $$ n=\sum^{k}_{i'=0}n_{i'}+k\;\;\;\;\;\;0\leq j'\leq k. $$ $n_{i'}$ is the number of steps (for walks $w$) in the ${\it G}_1$ coset which, once the renormalization-group map is applied, has the image $LX'_{i'}$. Thus, the total number of steps (for walks $w$) on the lattice is given by the sum of the steps within each L-block (${\it G}_1$ cosets) plus $k$ times 1 (due to the step out of the corresponding block).\ {\bf Theorem 1. The probability in eq(6), where $q(X_{i+1},X_{i})$ is chosen of the form $c|X_{i+1}-X_{i}|^{-\alpha}$ (c is a constant fixed up to normalization and $\alpha$ another constant), is a fixed point of the renormalization-group map $R(w)=w'$, provided $\beta =\alpha-d$}.\ {\bf Proof}. Using the definition of the renormalization-group map on the probability given in eq(6) and doing some elementary manipulations we arrive to the following expression \begin{equation} P'(w')=L^{\beta k}\;r^{k}\;e^{-rT} \prod^{k-1}_{j'=0}\;\;(q(LX'_{j'}- LX'_{j'+1})L^d)\times \end{equation} $$ \prod^{k}_{j'=0}\;\;\sum_{n_{j'}}\;\; r^{n_{j'}}\;\;(q_{1}(L^d-1))^{n_{j'}} \frac {(L^{\beta }t'_{j'})^{n_{j'}}} {n_{j'}!} $$ where $\prod^{n-1}_{i=0}q(X_{i+1}-X_{i})$ has been split into two factors; the first factor corresponding to jumps from one $L$-block to another $L$-block (different ${\it G}_1$ cosets) and the second factor corresponding to jumps inside the same $L$-block or ${\it G}_1$ coset. Function $q_1$ in eq(11) is the probability of jumping to a given point within the ${\it G}_1$ coset that has the image $LX'_{j'}$ (hereafter $\left({\it G}_{1}\right)_{X'_{j'}}$). There are $(L^d-1)$ possibilities with equal probability $q_1$ and $n_{j'}$ steps.\ {}From normalization, i.e. $\sum_{X\in {\it G}}q(X)=1$, we get $c=\frac {L^{\alpha -d}-1}{1-L^{-d}}$ and \newline $q_{1}(L^d-1)=1-L^{d-\alpha }$.\ On the other hand, we know that $T=L^{\beta }T'$, therefore, eq (11) becomes \begin{equation} P'(w')=L^{(d+\beta -\alpha)k}\;\;r^{k} \;\;\prod^{k-1}_{i'=0}q(X'_{i'+1}-X'_{i'})e^{-rL^{(d+\beta -\alpha)}T'} \end{equation} Provided $d+\beta -\alpha =0$, eq(12) is clearly a fixed point of the renormalization-group map $R$, i.e. $P'(w')=r^{k}\prod^{k-1}_{i'=0}q(X'_{i'+1}-X'_{i'})e^{-rT'}$. Q. E. D.\ Theorem 1 corresponds to the case worked out by Brydges, Evans, and Imbrie if we choose $\beta =2$, i.e. diffusive behavior \cite{Ev}.\ {\bf Theorem 2. If the probability in eq(6) where $q(X_{j},X_{j+1})$ is chosen of the form \begin{equation} q(X_j-X_{j+1})=c\left(\frac{1}{|X_j-X_{j+1}|^\alpha }+ \frac{1}{|X_j-X_{j+1}|^\gamma }\right) \end{equation} (c is a constant fixed up to normalization, $\alpha$ and $\gamma$ are constants, $\alpha\neq \gamma$) then $P(w)$ flows to the fixed point $P'(w')$ given in theorem 1, (i.e. $q(X_j-X_{j+1})$ is given as in theorem 1), under the renormalization-group map $R(w)=w'$, provided $\gamma >>\alpha$ (such that $log \left(\frac{L^{-\alpha}-L^{-\gamma}-2L^{d-\gamma-\alpha}}{L^{-\alpha}- 2L^{d-\gamma}}\right)\rightarrow 0$) and $\beta =d+\alpha $}. {\bf Proof}. Following the same ideas as in theorem 1, from normalization, we obtain $ c=\frac{(1-L^{d-\alpha})(1-L^{d-\gamma})} {(L^d-1)(L^{-\alpha}+L^{-\gamma}-2L^{d-\gamma -\alpha})} \;\;\mbox{ and } q_{1}=c(L^{-\gamma}+L^{-\alpha}) . $ Then, if $\gamma >> \alpha$, $P'(w')$ corresponds to \begin{equation} P'(w')=L^{(d+\beta -\alpha)k}\;\;r^{k} \;\;e^{-rL^{\beta }T'}\times \end{equation} $$ \times \prod^{k-1}_{j'=0}\frac {(L^{\alpha -d}-1)}{(1-L^{-d})} |X'_{j'}-X'_{j'+1}|^{-\alpha} e^{r(1-L^{d-\alpha})L^{\beta }T'} $$ If $d+\beta -\alpha=0$, eq(14) reduces to $P'(w')$ as given in Theorem 1.\ Q.E.D.\ Theorem 2 is presented here in order to learn about the L\'{e}vy process we are studying. This case is intended to answer the question about the feasibility of introducing perturbations in the probability and the possible results to be obtained. We are looking forward to studying of asymmetric and ``trapping" environments .\ \section{The renormalization-group map on \newline the weakly SARW} A configurational random walk model can be defined by assigning to every n-tuple of walks $w_1,...,w_n $ $ (n\geq 0)$ a statistical weight. For a simple random walk model, this is the product of the statistical weights for each of the n walks and can serve as a random walk representation of the Gaussian model \cite{Fer}\cite{Br}. The best known mathematical model that involves a self-repulsion term is the self-avoiding random walk. A self-avoiding walk of length n is a simple random walk which visits no site more than once. Unfortunately, it turns out that it is extremely difficult to obtain rigorous results from this model for $d\leq 4$ \cite{Ma}\cite{La1} \cite{La2}. However, there are other ways to include self-repulsion terms in random walk models \cite{La}. These split naturally into two categories: configurational measures where random walks are weighted by the number of (self-)intersections, and kinetically growing measures where random walks are produced from consistent measures that are generated by Markovian transition probabilities on the states space of simple random walks (these measures are non-Markovian on the state space of the lattice) \cite{La}. The weakly self-avoiding random (or Domb-Joyce) model and the Edwards model are examples of the first category \cite{La}. The ``true" (or ``myopic") self-avoiding walk and the Laplacian random walk are examples of the second category \cite{La}. In this paper we deal only with configurational measures. A simple random walk model can be thought of as being endowed with a configurational measure where the weight for the self-intersections of a walk (and/or among the n-tuple of walks $w_1,...,w_n $ $ (n\geq 0)$) is null. Configurational measures are measures on $\Lambda_n$, the space of simple random walks of length n. Let $P_U(w)$ be a probability on $\Lambda_n$ such that \begin{equation} P_U(w)=\frac{U(w)P(w)}{Z} \end{equation} where $Z=\int U(w)P(w)Dw$. In this paper we consider $P(w)$ as the probability on $\Lambda_n$ given in eq(6) and $U(w)$ as the interaction energy of the walks \cite{Fer}. Thus, to study the effect of the renormalization-group map on $P_U(w)$ we need to follow the trajectory of $U(w)$ after applying several times the renormalization-group map.\ Therefore, from the definition of the renormalization-group map in eq(8); $$ P'_{U'}(w')=L^{\beta k}\int P_{U}(w) \chi (R(w)=w')Dw $$ where $Z'=Z$, it follows that \begin{equation} U'(w')= \frac{\int Dw P(w)\chi (R(w)=w')U(w)} {\int Dw P(w)\chi (R(w)=w')} \end{equation} Note that eq(16) can be viewed as the expectation of $U(w)$ given that the renormalization-group map is imposed, calculated using $P(w)$ on $\Lambda_{n}$ defined as in eq(6). Therefore, and hereafter, to simplify notation, we write eq(16) as $U'(w')=< U(w) >_{w'}$.\ In this Section we deal with $U(w)$ factorizable in terms of the interaction energy with null weight for the (self-)intersection of $n$ $(n=2,3,...,etc.)$ random walks (i.e. a simple random walk factor), and the interaction energies that weight the intersection of n random walks. Hereafter, as a hypothesis, we assume all the factors in $U(w)$ (functions of $w$) as independent, simple, random variables. Thus, the conditional (given the $R(w)=w'$ map) expectation of $U(w)$ is the product of conditional (given the $R(w)=w'$ map) expectation of each factor in $U(w)$. This hypothesis follows the same spirit as in the approach used in polymer networks \cite{Du} and can be seen as a consequence of factorizability and locality of the map. See Figure 2.\ To start with, we study the simple random walk model such that \begin{equation} U(w)=\prod_{X\in {\it G}} e^{-a\sum _{j\in J_{X}}\;\;t_j}, \end{equation} where $J_{X}=\{j\in \{0,...,n\}|X_j=X \}$ for $w(t_0+,...,+t_j)=X_j$ and $X\in {\it G}$.\ {\bf Theorem 3. The probability $P_U(w)$ for the simple random walk model where U(w) is given by eq(17), is a fixed form of the \newline renormalization-group map $R(w)=w'$ such that, after applying the renormalization-group map, $a'=L^{\beta }a$}.\ {\bf Proof}. Let us split the product on sites in the lattice in $ \prod_{X\in {\it G}} e^{-a\sum _{j\in J_{X}}\;\;t_j} $ into two parts. The first one, i.e. $ \prod_{X'_{i'}\in {\it G}/{\it G_1}} \mbox{ corresponds to } \prod_{X'_{i'}\in {\it G}} $ due to the hierarchical structure of the lattice. The second one, i.e. $ \prod_{X_{i}\in ({\it G_1})_{X'_{i'}}} $ stands for the L-block (${\it G_1}$ coset) of the lattice ${\it G}$ that, under the renormalization-group transformation, maps to $LX'_{i'},\;\;(0\leq i'\leq k)$. There are $k$ replicas of this. If we again split $\prod^{n-1}_{i=0}q(X_{i+1}-X_i)$ into two factors, as was done in theorem 1. We obtain \begin{equation} U'(w')=\prod_{X'_{i'}\in {\it G}} \left\{ \prod_{j'\in J_{X'_{i'}}} e^{q_{1}(L^d-1)r(L^{\beta}t'_{j'})} \right\}^{-1}\times \end{equation} $$ \left\{ \sum_{n_{i'}}\;\int\prod_{i\in I_{X'_{i'}}} \;dt_{i}\;\prod_{j'\in J_{X'_{i'}}} \;\delta (\sum^{m_{j'}}_{{i}=m_{j'-1}+1}\;\;t_{i}- L^{\beta }t'_{j'})\times \right. $$ $$ \sum_{X_{i}\in ({\it G}_1)_{X'_{i'}}} \prod^{m_{j'}}_{{i}=m_{j'-1}+1}\;\;\; \chi (X_{i}\in LX'_{j'})\;\; (q_{1}(L^d-1)r)^{n_{i'}} \left. e^{-a\sum_{\left(X_{i}\in {\it G_1}\right)_{X'_{i'}}} \sum _{j\in J_{X_{i}}}\;\; t_{j}} \right\} $$ where we have defined , for $X_{i}\in w$ and $X'_{i'}\in w'$; \begin{equation} I_{X'_{i'}}=\left\{i\;|\;X_{i}\in LX'_{i'}\right\}= \end{equation} $$ \bigcup_{j'\in J_{X'_{i'}}} \left\{i\;|\;m_{j'-1}+1\leq i\leq m_{j'}\right\}= \bigcup_{j'\in J_{X'_{i'}}} \left\{i\;|\; 0\leq i\leq n_{j'}\right\}. $$ Rearranging the double sum in the exponential of eq(18), we obtain \begin{equation} U'(w')=\prod_{X'\in {\it G}'} e^{-aL^\beta \sum_{j'\in J_{X'}}t'_{j'}}. \end{equation} Q.E.D.\ The next model we want to study is a weakly self-avoiding random walk (or Domb-Joyce model), with a configurational measure in which double intersections of walks are penalized by (roughly speaking) a factor of $e^{-\lambda}$. As $\lambda \rightarrow \infty$, this reduces to random walks with strict mutual avoidance. Recall that this weakly model (with $\lambda >0$) and the perfect self-avoiding random walk, are in the same universality class (this implies that the critical exponents are the same). If $\lambda=0$, this corresponds to a simple random walk model. Next Theorem is an important result of this paper; it is a random walk version of the field theoretical approach \cite{Ev} with improved physical intuition and the key stone in our method to obtain the end-to-end distance of the weakly SARW. The renormalization-group map applied on a weakly SARW involves the paramaters $\gamma_1$, $\gamma_2$, $\beta_1$, $\eta$, $A$, $B$, $C$. These are some conditional expectations of local times for different topologies in both, $w$ and $w'$ random walks and are precisely defined in Figure 3.\ In Figure 3, $\gamma_1$ and $\gamma_2$ correspond to $O(\lambda)$ and $O(\lambda^2)$ contributions in Figure 2.b), respectively. $\beta_1$ corresponds to $O(\lambda^2)$ contribution in figure 2.c) and $\eta$ corresponds to $O(\lambda^2)$ contribution in Figure 2.d). Although $A$, $B$, and $C$ are not depicted in Figure 2, it is straightforward to figure out the corresponding pictures.\ {\bf Theorem 4. For the weakly SARW with interaction energy} \begin{equation} U(w)=\prod_{X_i\in {\it G}} e^{-\xi\sum_{i\in J_{X_{i}}}t_{i}-\lambda \sum_{i<j\in J_{X_{i}}} t_{i}t_{j} {\bf 1}_{\left\{w(t_i)=w(t_j)\right\} }} \end{equation} {\bf $\xi_{2}<0$ and $\lambda >0$ being (small) constants, the probability $P_U(w)$ flows to a fixed form after the renormalization-group transformation is applied. This fixed form is characterized by the interaction energy} \begin{equation} U'(w')= \prod_{X'_{i'}\in{\it G}} e^{ -\xi' \sum_{i_{\alpha_{1}}'\in J_{X'_{i'}}} t'_{i_{\alpha_{1}}'}- \lambda' \sum_ {\stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}))}} \times \end{equation} $$ \left\{1+\eta'_{1} \sum_{ \stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'< k_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}', k_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} t'_{k_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}) =w(t'_{k'_{\alpha_{1}}}))}+ \right. $$ $$ +\left. \eta'_{2} \sum_ {\stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}))} +\eta'_{3}\sum_{i_{\alpha_{1}}'\in J_{X'_{i'}}} t'_{i_{\alpha_{1}}'}\right\}+r'. $$ {\bf Here} \begin{eqnarray} \xi' & = & L^{\beta}\xi+\xi'_2 \\ \xi'_{2} & = & \gamma _{1}\lambda-\gamma _{2}\lambda^2+O(\lambda^3)\\ \lambda' & = & L^{(2\beta-d)}\lambda-\beta_{1}\lambda^{2}+O(\lambda^3) \\ \eta'_{1} & = & \eta_{1}A+\eta \lambda^2 \\ \eta'_2 & = & \eta_1B+L^{(2\beta-d)}\eta_2 \\ \eta'_3 & = & \eta_1C+\eta_2\gamma_{1}+L^{\beta}\eta_{3} \\ r'& \sim & O\left(\lambda^3\right). \end{eqnarray} {\bf Proof}. From our initial hypothesis, follows \begin{equation} U'(w')= \left\langle U_{\xi}(w)\right\rangle_{w'} \left\langle U_{\lambda}(w)\right\rangle_{w'}, \end{equation} to errors $O(\xi\lambda)$ at each site. Here we introduce $\xi=O(\lambda^2)$, so this can be put in $r$. Besides the subindices $\xi$ and $\lambda $ identify the two factors in eq(21). We know, from theorem 3, the trajectory of $\left\langle U_{\xi}\left(w\right)\right\rangle_{w'}$, so we only have to study the trajectory of $\left\langle U_{\lambda}\left(w\right)\right\rangle_{w'}$.\ We split the product in the interaction term $\left\langle U_{\lambda}\left(w\right)\right\rangle_{w'}$ into two parts and identify $ \prod_{X'_{i'}\in {\it G}/{\it G_1}} \mbox{ with } \prod_{X'_{i'}\in {\it G}} $ , therefore; \begin{equation} \left\langle U_{\lambda}\left(w\right)\right\rangle_{w'} =\prod_{X'_{i'}\in {\it G}} \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}}}_{w'} \;\mbox{ where } \end{equation} \begin{equation} \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}}}_{w'}= \left\{ \prod_{j'\in J_{X'_{i'}}} e^{q_{1}(L^d-1)r(L^{\beta}t'_{j'})} \right\}^{-1}\times \end{equation} $$ \left\{ \sum_{X_{i}\in ({\it G}_{1})_{X'_{i'}}}\; \sum_{n_{i'}}\int\prod_{i\in I_{X'_{i'}}}dt_{i} \right. \prod_{j'\in J_{X'_{i'}}} \;\delta (\sum^{m_{j'}}_{{i}=m_{j'-1}+1}\;\;t_{i}-L^{\beta }t'_{j'}) \prod^{m_{j'}}_{{i}=m_{j'-1}+1} \chi (X_{i}\in LX'_{j'}) $$ $$ \times \prod^{n-1}_{i=0} q(X_{i+1}-X_i) \left. r^{n_{i'}}\; e^{-\lambda\sum_{X_{i}\in ({\it G_1})_{X'_{i'}}} \sum _{i_{\alpha_{1}}< j_{\alpha_{1}}\in J_{X_{i}}}\;\; t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))}} \right\} $$ Recall that $w'$ is fixed; i.e. the walk, after the renormalization-group map is applied, visits each site $X'_{i'}\in {\it G}$ a fixed number of times $n'^{*}_{i'}$. Let us assume $w'$ is such that $\left\{n'^{*}_{i'}\right\}>0$, and at least once $n'^{*}_{i'}=1,2,3$ on ${\it G}$. See example in Figure 2. We ask for this condition to hold, in order to learn about the flow of the interaction constant in the (self-) intersection of 2 and 3 random walks. In other words, we ask for a fixed and not totally arbitrary topology for the renormalized random walk $w'$ on ${\it G}$. To make this condition explicit we rewrite eq(31) as; \begin{equation} \left\langle U_{\lambda}\left(w\right)\right\rangle_{w'} =\prod_{X'_{i'}\in {\it G}} \prod_{\left(n'^{*}_{i'}\right)^{k}_{i'=0}} \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}}} _{w'_{n'^{*}_{i'}}} \end{equation} Here, $ \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}}} _{w'_{n'^{*}_{i'}}} $ is the renormalized interaction energy for all possible topologies of walks $w$ inside the ${\left({\it G}_{1}\right)_{X'_{i'}}}$ coset that, once the renormalization-group map is applied, corresponds to a fixed topology in $LX'_{i'}\in {\it G}$.\ Let us introduce a formal Taylor series expansion in $\lambda$, then; \begin{equation} \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}}} _{w'_{n'^{*}_{i'}}}= \sum^{\infty}_{s=1}\frac{\left(-\lambda \right)^s}{s!} \end{equation} $$ \left\langle \sum_{X_i\in {({\it G}_1)}_{X'_{i'}}}\; \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}<;...; <i_{\alpha_{s}}< j_{\alpha_{s}}} {\left\{i_{\alpha_{1}},...,i_{\alpha_{s}} ; j_{\alpha_{1}},...,j_{\alpha_{s}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))} \times ... \right. $$ $$ \left. ...\times t_{i_{\alpha_{s}}}t_{j_{\alpha_{s}}} {\bf 1}_{(w(t_{i_{\alpha_{s}}})=w(t_{j_{\alpha_{s}}}))} \chi\left(\left(i_{\alpha_{1}},...,i_{\alpha_{s}}; j_{\alpha_{1}},...,j_{\alpha_{s}}\right)\in (i'_{\alpha_{1}},..., i'_{\alpha_{n'^{*}_{i'}}}) \right) \right\rangle ^{l.c.}_{w'} $$ In this formal series expansion, we are writing $ \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}}} _{w'_{n'^{*}_{i'}}} $ in terms af all possible classes of topology for walks $w$ inside the ${\left({\it G}_{1}\right)_{X'_{i'}}}$ coset. Each class is an element of this Taylor series and corresponds to a fix number $s$ of double (self)-intersections, weighted by $\lambda^s$. Here, the superscript $l.c.$ means linear contribution. We take into account only linear contributions to conditional expectations. This approach is considered to avoid double-counting sites in walks.\ To start with, we analyze explicitly the case $n'^{*}_{i'}=1$, (for some $0\leq i'\leq k$) up to 2nd order in $\lambda$; \begin{equation} \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}}} _{w'_{n'^{*}_{i'}=1}}= 1-\lambda\times \end{equation} $$ \left\langle \sum_{X_i\in {({\it G}_1)}_{X'_{i'}}}\; \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}} {\left\{i_{\alpha_{1}}, j_{\alpha_{1}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} \left. {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))} \chi\left(\left(i_{\alpha_{1}}, j_{\alpha_{1}}\right)\in i'_{\alpha{1}} \right) \right\rangle ^{l.c.}_{w'} \right. $$ $$ +\frac{\lambda^{2}} {2} \left\langle \sum_{\left(X_{i}\in {\it G}_{1}\right)_{X'_{i'}}}\; \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< i_{\alpha_{2}}< j_{\alpha_{2}}} {\left\{i_{\alpha_{1}},i_{\alpha_{2}}, j_{\alpha_{1}},j_{\alpha_{2}}\right\} \in J_{\left(X_{i}\right)}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}}\times \right. $$ $$ \times\left. {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))} t_{i_{\alpha_{2}}}t_{j_{\alpha_{2}}} {\bf 1}_{(w(t_{i_{\alpha_{2}}})=w(t_{j_{\alpha_{2}}}))} \chi\left(\left(i_{\alpha_{1}},i_{\alpha_{2}}, j_{\alpha_{1}},j_{\alpha_{2}}\right)\in i'_{\alpha_{1}} \right) \right\rangle ^{l.c.}_{w'}+ r'_{1}, $$ where $r'_1\sim O(\lambda^3)$.\ Thus, eq(35) is written as \begin{equation} \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}},l.c.} _{w'_{n'^{*}_{i'}=1}}= 1-\left(\gamma_{1}\lambda-\gamma_{2}\lambda^{2} +O(\lambda^3)\right) \sum_{i_{\alpha_{1}}'\in J_{X'_{i'}}} t'_{i_{\alpha_{1}}'} \end{equation} $$ \cong e^{-\xi'_{2} \sum_{i_{\alpha_{1}}'\in J_{X'_{i'}}} t'_{i_{\alpha_{1}}'}}. $$ where $\xi'_{2}$ is given as in eq(24), \begin{equation} \gamma_{1}= \left\langle \sum_{X_i\in ({\it G}_1)_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))} \chi\left(\left(i_{\alpha_{1}},j_{\alpha_{1}}\right)\in i'_{\alpha_{1}} \right) \right\rangle ^{l.c.}_{w'}\;\mbox{ and} \end{equation} \begin{equation} \gamma_{2}= \frac{1}{2} \left\langle \sum_{X_i\in ({\it G}_1)_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< i_{\alpha_{2}}< j_{\alpha_{2}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}, i_{\alpha_{2}},j_{\alpha_{2}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))}\times \right. \end{equation} $$ \left.\times t_{i_{\alpha_{2}}}t_{j_{\alpha_{2}}} {\bf 1}_{(w(t_{i_{\alpha_{2}}})=w(t_{j_{\alpha_{2}}}))} \chi\left(\left(i_{\alpha_{1}},j_{\alpha_{1}}, i_{\alpha_{2}},j_{\alpha_{2}}\right)\in i'_{\alpha_{1}} \right) \right\rangle ^{l.c.}_{w'}. $$ $\gamma_1$ and $\gamma_2$ are explained in Figure 3.\ In the same spirit we study the case $n'^{\ast}_{i'}=2$, (for some $0\leq i'\leq k$) up to second order in $\lambda $; \begin{equation} \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}}lc} _{w'_{n'^{*}_{i'}=2}}=1-\left( L^{(2\beta-d)}\lambda-\beta_{1}\lambda^{2}+O(\lambda^3)\right)\times \end{equation} $$ \times \sum_ {\stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}))} $$ $$ \cong e^{- \lambda' \sum_ {\stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}))}}. $$ where $\lambda'$ is given by eq(25) and \begin{equation} \beta_1=\frac{1}{2} \left\langle \sum_{X_i\in ({\it G}_1)_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< i_{\alpha_{2}}< j_{\alpha_{2}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}, i_{\alpha_{2}},j_{\alpha_{2}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))} \right. \times \end{equation} $$ \times\left. t_{i_{\alpha_{2}}}t_{j_{\alpha_{2}}} {\bf 1}_{(w(t_{i_{\alpha_{2}}})=w(t_{j_{\alpha_{2}}}))} \chi\left(\left(i_{\alpha_{1}},j_{\alpha_{1}}, i_{\alpha_{2}},j_{\alpha_{2}}\right)\in \left(i'_{\alpha_{1}}< j'_{\alpha_{1}}\right) \right) \right\rangle ^{l.c.}_{w'}. $$ $\beta_1$ is explained in Figure 3.\ Finally, let us present the factor $n'^{*}_{i'}=3$, (for some $0\leq i'\leq k$), up to 2nd order in $\lambda$; \begin{equation} \left\langle U_{\lambda}\left(w\right) \right\rangle^{\left({\it G}_{1}\right)_{X'_{i'}},l.c.} _{w'_{n'^{*}_{i'}=3}}= \left(1+ \right. \end{equation} $$ +\eta\lambda^{2} \sum_{ \stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'< k_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}', k_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} t'_{k_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}) =w(t'_{k'_{\alpha_{1}}}))} \left. +r'_{3}\right) $$ where $\eta$ is \begin{equation} \eta=\frac{1}{2} \left\langle \sum_{X_{i}\in ({\it G}_{1})_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< i_{\alpha_{2}}< j_{\alpha_{2}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}, i_{\alpha_{2}},j_{\alpha_{2}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))} \right. \times \end{equation} $$ \times\left. t_{i_{\alpha_{2}}}t_{j_{\alpha_{2}}} {\bf 1}_{(w(t_{i_{\alpha_{2}}})=w(t_{j_{\alpha_{2}}}))} \chi\left(\left(i_{\alpha_{1}},j_{\alpha_{1}}, i_{\alpha_{2}},j_{\alpha_{2}}\right)\in \left(i'_{\alpha_{1}}< j'_{\alpha_{1}}<k'_{\alpha_{1}}\right) \right) \right\rangle ^{l.c.}_{w'} . $$ $\eta$ is explained in Figure 3.\ Note that in eq(41) {\footnotesize \begin{equation} \left\langle \sum_{X_{i}\in ({\it G}_{1})_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}} {\left\{i_{\alpha_{1}}, j_{\alpha_{1}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))} \chi\left(\left(i_{\alpha_{1}}, j_{\alpha_{1}}\right)\in \left(i'_{\alpha_{1}}< j'_{\alpha_{1}} < k'_{\alpha_{1}}\right)\right) \right\rangle ^{l.c.}_{w'}=0 \end{equation}} for all set of walks $\{w\}\in \Gamma_{n}$, being $\Gamma_n$ the space of random walks with only double (self-)intersecting walks.\ Note that the case $n'^{*}_{i'}=1$, for some $0\leq i'\leq k$, shows how the weakly (self-)avoiding random walk on the $\left({\it G}_{1}\right)_{X'_{i'}}$ coset can lead to a ``mass" term; i.e. a local time contribution, after the transformation is applied. This shall be used in next section to obtain the asymptotic end-to-end distance of a weakly SARW on a hierarchical lattice, $d=4$, heuristically. Similarly, the case $n'^{*}_{j'}=3$, for some $j'\neq i'$ and $0\leq j'\leq k$, shows how a different realization on the ${\it G}_{1}$ cosets of the weakly SARW can render, a pair of double intersections of random walks inside the $\left({\it G}_{1}\right) _{X'_{j'}}$ coset, to the triple intersection of renormalized walks in $X'_{j'}$ after the renormalization-group map is applied. See Figure 2.\ Writing together cases $n'^{*}_{i'}=1,2,3$ we obtain \begin{equation} U'(w')= \prod_{X'_{i'}\in{\it G}} \end{equation} $$ e^{ -\left(L^{\beta}\xi+\xi'_{2}\right) \sum_{i_{\alpha_{1}}'\in J_{X'_{i'}}} t'_{i_{\alpha_{1}}'}- \lambda' \sum_ {\stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1} } }) =w(t'_{j'_{\alpha_{1}} })) } }\times $$ $$ \left\{1+\eta\lambda^{2} \sum_{ \stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'< k_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}', k_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} t'_{k_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}) =w(t'_{k'_{\alpha_{1}}}))} \right\}+r'. $$ where $r'\sim O(\lambda^{3})$ bounds the cosets where $n'^{\ast}_{i'}\geq 4$, for some $0\leq i'\leq k$. Note that $\lambda^3$ is the leading contribution to $r'$ if $n'^{\ast }_{i'}= 4$. As $n'^{\ast }_{i'}$ increases the leading contribution decreases. Here $$ \xi'=\xi'_{2}+L^{\beta}\xi $$ $$ \eta'_1=\eta\lambda^2\;\;\;\;\;\;\; \lambda'=L^{(2\beta-d)}\lambda-\beta'\lambda^2+O(\lambda^3) $$ We apply once more the renormalization-group map to eq(44). In sake of clarity let us supress primes in eq(44), so the primed terms always correspond to the renormalized ones. From the initial hypothesis, approaching the three walks intersection factor in eq(44) to an exponential, factorize this as done in eq(33), then expanding the exponential up to the first order, i.e. $\sim \left(\eta\lambda^2\right)$; from the result in theorem 3, eq(36), eq(39) and eq(41), follows \begin{equation} U'(w')= \end{equation} $$ \prod_{X'_{i'}\in{\it G}'} e^{ -\xi' \sum_{i_{\alpha_{1}}'\in J_{X'_{i'}}} t'_{i_{\alpha_{1}}'}- \lambda' \sum_ {\stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}}) =w(t'_{j'_{\alpha_{1}}}))}}\times $$ $$ \left\{1+\eta'_{1} \sum_{ \stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'< k_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}', k_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} t'_{k_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}) =w(t'_{k'_{\alpha_{1}}}))} \right. $$ $$ +\left. \eta'_{2} \sum_ {\stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}}) =w(t'_{j'_{\alpha_{1}}}))} +\eta'_{3}\sum_{i_{\alpha_{1}}'\in J_{X'_{i'}}} t'_{i_{\alpha_{1}}'}\right\}+r'. $$ where $r'\sim O(\lambda^3)$ and $$ \xi'_{2}= \gamma_{1}\lambda-\gamma_{2}\lambda^2 +O(\lambda^3), $$ $$ \xi'=L^{\beta}\xi+\xi'_2, $$ $$ \lambda'=L^{(2\beta -d)}\lambda-\beta_{1}\lambda^2+ O(\lambda^3); \;\;\; \eta'_{1}=\eta_{1}A+\eta\lambda^{2} $$ $$ \eta'_{2}=\eta_{1}B, \;\;\;\; \eta'_{3}=\eta_{1}C $$ In eq(45) $A$, $B$ and $C$ are respectively given as follows; \begin{equation} A=\left\langle \sum_{X_i\in ({\it G}_1)_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< k_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}},k_{\alpha_{1}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} t_{k_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}) =w(t_{k_{\alpha_{1}}}))} \right. \times \end{equation} $$ \times\left. \chi\left(\left(i_{\alpha_{1}},j_{\alpha_{1}},k_{\alpha_{1}}\right) \in \left(i'_{\alpha_{1}},j'_{\alpha_{1}},k'_{\alpha_{1}}\right) \right) \right\rangle ^{l.c.}_{w'}, $$ \begin{equation} B=\left\langle \sum_{X_{i}\in ({\it G}_{1})_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< k_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}},k_{\alpha_{1}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} t_{k_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}) =w(t_{k_{\alpha_{1}}}))} \right. \times \end{equation} $$ \times\left. \chi\left(\left(i_{\alpha_{1}},j_{\alpha_{1}},k_{\alpha_{1}}\right) \in \left(i'_{\alpha_{1}}<j'_{\alpha_{1}}\right) \right) \right\rangle ^{l.c.}_{w'} \;\;{\mbox and} $$ \begin{equation} C=\left\langle \sum_{X_{i}\in ({\it G}_{1})_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}<j_{\alpha_{1}}< k_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}},k_{\alpha_{1}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} t_{k_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}) =w(t_{k_{\alpha_{1}}}))} \right. \times \end{equation} $$ \times\left. \chi\left(\left(i_{\alpha_{1}},j_{\alpha_{1}},k_{\alpha_{1}}\right) \in \left(i'_{\alpha_{1}}\right) \right) \right\rangle ^{l.c.}_{w'}. $$ $A$, $B$ and $C$ are explained in Figure 3.\ Note that, to get the result in eq(45), we have studied the trajectory, under the renormalization-group map R, of \begin{equation} \left\{ \prod_{X_{i}\in ({\it G}_{1})_{X'_{i'}}}\; e^{ \eta'_1 \sum_{ \stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< k_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}, k_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} t_{k_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}) =w(t_{k_{\alpha_{1}}}))} } \right\}. \end{equation} When we apply the renormalization-group map to eq(49) we end up with the contributions in $A,B$ and $C$. In other words, we use the same procedure above developed for the double (self-)intersection of random walks but for the triple (self-)intersection of random walks up to order $\lambda^{2}$, this corresponds to the first term in the corresponding Taylor series expansion.\ Applying the renormalization-group map to eq(45) we obtain eq(22) (eq(44) and eq(45) are particular cases of eq(22)). Then, the proof follows from induction. We apply the renormalization-group map to $U'(w')$ (eq(22)). Recall that we supress primes in eq(22), thus primed terms are the renormalized ones. From our original hypotheses and due to the hierarchical structure of the lattice; \begin{equation} U'(w')= \end{equation} $$ \prod_{X'_{i'}\in{\it G}} \left\langle \prod_{X_{i}\in ({\it G_1})_{X'_{i'}}} e^{ -\xi \sum_{i_{\alpha_{1}}\in J_{X_{i}}} t_{i_{\alpha_{1}}}} \right\rangle^{l.c}_{w'}\times $$\ $$ \times \left\langle \prod_{X_{i}\in ({\it G_1})_{X'_{i'}}} e^{ -\lambda \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))}} \right\rangle^{l.c}_{w'}\times $$\ $$ \left\langle1+\eta_{1} \sum_{X_{i}\in ({\it G_1})_{X'_{i'}}} \sum_{ \stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< k_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}, k_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} t_{k_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}) =w(t_{k_{\alpha_{1}}}))}+ \right. $$\ $$ +\eta_{2} \sum_{X_{i}\in ({\it G_1})_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))}+ $$ $$ \left. +\eta_{3} \sum_{X_{i}\in ({\it G_1})_{X'_{i'}}} \sum_{i_{\alpha_{1}}\in J_{X_{i}}} t_{i_{\alpha_{1}}}\right\rangle^{l.c}_{w'} +r'. $$ Assuming ${n^{\ast}}'_{i'}=1,2,3$ and carrying out a Taylor series expansion up to order $\lambda^{2}$, it is straightforward to prove that eq(50) leads to eq(22), provided we use the same bookkeeping device above explained. We just need to apply theorem 3, eq(36), eq(39), eq(41) and \begin{equation} \left\langle \prod_{X_{i}\in ({\it G_1})_{X'_{i'}}} exp( \eta_{1} \sum_{ \stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< k_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}, k_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} t_{k_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}) =w(t_{k_{\alpha_{1}}}))}+ \right. \end{equation} $$ \left.+ \eta_{2} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))} +\eta_{3} \sum_{i_{\alpha_{1}}\in J_{X_{i}}} t_{i_{\alpha_{1}}})\right\rangle^{l.c}_{w'}= $$ $$ \prod_{\left(n'^{*}_{i'}\right)^{n'}_{i'=0}} \left\langle \prod_{X_{i}\in ({\it G_1})_{X'_{i'}}} e^{ \eta_{3} \sum_{i_{\alpha_{1}}\in J_{X_{i}}} t_{i_{\alpha_{1}}}} \right\rangle^{l.c.}_{w'_{n'^{*}_{i'}}} $$ $$ \times \left\langle \prod_{X_{i}\in ({\it G_1})_{X'_{i'}}} e^{\eta_{1} \sum_{ \stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< k_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}, k_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} t_{k_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}) =w(t_{k_{\alpha_{1}}}))}} \right\rangle^{l.c}_{w'_{{n^{*}}'_{i'}}} $$ $$ \times \left\langle \prod_{X_{i}\in ({\it G_1})_{X'_{i'}}} e^{ \eta_{2} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))}} \right\rangle^{l.c.}_{w'_{{n^{*}}'_{i'}}} = $$ $$ \left\{1+A\eta_{1} \sum_{ \stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'< k_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}', k_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} t'_{k_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}) =w(t'_{k'_{\alpha_{1}}}))} + \right. $$ $$ +\eta'_2 \sum_ {\stackrel{i_{\alpha_{1}}'< j_{\alpha_{1}}'} {\left\{i_{\alpha_{1}}',j_{\alpha_{1}}'\right\}\in J_{X'_{i'}}}} t'_{i_{\alpha_{1}}'}t'_{j_{\alpha_{1}}'} {\bf 1}_{(w(t'_{i'_{\alpha_{1}}})=w(t'_{j'_{\alpha_{1}}}))} \left. +\eta'_{3}\sum_{i_{\alpha_{1}}'\in J_{X'_{i'}}} t'_{i_{\alpha_{1}}'}\right\}+r'_{3} $$ where $A$, $\eta'_{2}$ and $\eta'_{3}$ are given by Figure 3, eq(27), and eq(28); $r'\sim O(\lambda^{3})$. Note that $A\eta_1$ in eq(51) plus the contribution from eq(41) defines $\eta'_{1}$ as done in eq(26). Eq(51) is written up to $O(\lambda^2)$ terms.\ Q.E.D.\ We claim theorem 4 is the space-time renormalization-group trajectory of the weakly SARW energy interaction studied by Brydges, Evans and Imbrie \cite{Ev}, provided $\beta=2$ and $d=4$. In reference [5] the trajectory of a $\lambda\phi^{4}$ superalgebra valued interaction was studied (this can be understood in terms of intersection of random walks due to the Mc Kane, Parisi, Sourlas Theorem \cite{Mk}) using a field-theoretical version of the renormalization-group map. The field theory is defined on the same hierarchical lattice we are studying here. In this paper, we provide exact probabilistic expressions for $\lambda'$ and $\xi'$ (which are not given in reference [5]), these are crucial to propose an heuristic proof for the asymptotic behavior of the end-to-end distance of a weakly SARW. To do so we just need to calculate $\gamma_{1}$ from eq (37) and $\beta_{1}$ from eq(40).\ Finally, we summarize important features of our method; \newline a) the conditional expectation of $U(w)$ can be approached in terms of the product of conditional expectations.\ b) We take into account only linear contributions to conditional expectations for probabilities on $\Lambda_{n}$.\ c) Formal Taylor series expansions are introduced.\ d) We assume our model to be such that each step the renormalization-group map is applied, the number of times the renormalized walk visits any site of the lattice is 1, 2 and 3 at least once (i.e. a fixed and not totally arbitrary topology for the renormalized random walk). See Figure 2. e) Finally, we take advantage of the hierarchical structure of the lattice. Since ${\it G}'={\it G}/{\it G}_{1}\approx {\it G}$ and the map is local, the renormalization-group transformation descends to the study of walks in the ${\it G}_1$ cosets.\ {}From all of these, we obtain, after applying the renormalization-group map, the fixed form for a weakly SARW that penalizes, roughly speaking, the (self-)intersection of two random walks by a factor ($e^{-\lambda}$, $\lambda >0$ and small). Furthermore, this fixed form is the random walk version (for $d=4$, $\beta=2$) of the one obtained from a field-theoretical renormalization-group map for a $\lambda \phi^{4}$ model recently reported by Brydges, Evans and Imbrie \cite{Ev}. We obtain an exact probabilistic expression for the parameters that appear in the flow of the interaction factor $\lambda$ which is not given in reference [5]. This shall be used in next section for the heuristic study of the asymptotic behavior of the end-to-end distance for a weakly SARW model that punishes the (self-)intersection of two random walks. \section{Asymptotic end-to-end distance of a weakly SARW on a hierarchical lattice in dimension four. An heuristic example as a testing ground.} The process of renormalizing the lattice is completed by reducing all dimensions of the new lattice by a factor $L$ each step the renormalization-group map is applied so we end up with exactly the same lattice we start with. For a diffusive simple random walk model we reduce waiting times by $L^{2}$ each step we apply the renormalization-group transformation. Moreover, when we iterate probabilities, the end-to-end distance shrinks by a factor $L$ at each interaction, because in renormalizing the lattice we divide every length, including the end-to-end distance, by $L$ \cite{Ma} \cite{Bi}. {}From this viewpoint we intend to understand, heuristically, the asymptotic end-to-end distance of a weakly SARW on a hierarchical lattice in $d=4$, thereby providing a new probabilistic meaning to this magnitude.\ For weakly SARW, we generalize the standard scaling factor for local times of the renormalization transformation above as described by including, up to $O(\lambda)$, the contribution of the self-repulsion term to renormalized local times. Namely, from the renormalization-group map on weakly SARW, renormalized local times are generated from the interaction. In the field theoretical approach this corresponds to generating mass. Equivalently, we can say that the interaction kills the process at a specific rate. If we take into account only $O(\lambda)$ contributions to this and follow standard thinking, the well known asymptotic end-to-end distance for the weakly SARW in $d=4$ follows. By including higher order contributions in $\lambda$ to renormalized local times (as we have already shown this is not the case for weakly SARW on the hierarchical lattice, because these contributions are no significant), and/or different dimension for the lattice, the functional form of the end-to-end distance changes drastically. Moreover, from our method, the exponent of the logarithmic correction involved is expressed in terms of conditional expectations for random walks on the lattice, that upon calculation, give the well known exponent. In Figure 4 the contributions to the scaling factor proposed for the weakly SARW used to explain the asymptotic end-to-end distance in the hierarchical lattice are depicted.\ We remark that the proposition presented in this section involves heuristic considerations in order to understand, from a probabilistic real-space viewpoint, the asymptotic end-to-end distance for the weakly SARW on a hierarchical lattice in $d=4$. This has already been conjectured heuristically before by other means. Recently a rigorous proof has been given, provided properties of the Green function are known, in the field theoretical approach \cite{Im}. Our proposition is anyway presented as a testing ground for our method, and for giving probabilistic meaning to the exponent involved in the logarithmic correction. Once the method shows to be useful for explaining well known results (at least heurstically), we shall apply this on more complicated cases, for example kinetically growing measure model. These are renormalizable, in the field theoretical limit, only for particular cases. In the process of taking the continuum limit of these discrete models some memory is lost. Our method is suitable of being applied on the discrete models. This can be done both, heuristically and rigorously.\ A final remark before introducing the main point of this section is about the finiteness of moments for random walks on a hierarchical lattice. The end-to-end distance for the weakly SARW, d=4, on the hierarchical lattice, is independent of the moment used to obtain it, as should be, provided this is finite. Let $\left\langle w^{\alpha}(T)\right\rangle$ be an $\alpha$-moment of the random walk, it is known that the only finite moments for diffusive random walks on a hierarchical lattice are $0<\alpha<2$ \cite{Im}. This range of $\alpha$ values is used to obtain the end-to-end distance in the following \newline {\bf Proposition. For d=4, up to $O(\lambda)$, the generated renormalized local times (mass for the field or killing rate for the process), from applying the renormalization-group map on the interaction, is such that the asymptotic behavior of the end-to-end distance for a weakly SARW that penalizes the intersection of two random walks is $T^{1/2}log^{1/8}T$ as T tends to infinity}.\ {\bf Proof}. After applying $(p)$ times the renormalization-group transformation on $\left\langle w^{\alpha}(T)\right\rangle^{1/\alpha}$ we have \begin{equation} \left\langle w^{\alpha}(T)\right\rangle^{1/\alpha}= \frac{\left\langle w^{\alpha}(1)\right\rangle^{1/\alpha (0)}}{L^{p}} \end{equation} where we have chosen a system of units such that, for $p=0$, $T=1$. Hereafter, $\left\langle w^{\alpha}(1)\right\rangle^{1/\alpha (0)}$=D, constant. Here, we are following the standard procedure for scaling length type magnitudes \cite{Bi} ( i.e. $\left\langle w^{\alpha}(T)\right\rangle^{1/\alpha'}$= $\frac{\left\langle w^{\alpha}(T)\right\rangle^{1/\alpha}}{L}$). Moreover, by $\left\langle w^{\alpha}(1)\right\rangle^{1/\alpha}$ we mean $\left\langle w^{\alpha}(1)\right\rangle^{1/\alpha (p)}$. Since in renormalizing the lattice we divide every length, including the end-to-end distance by $L$, then, upon p iterations, eq(52) follows. This is exactly what is done in scaling correlation lengths but used here on the end-to-end distance, both length type magnitudes. So eq(52) becomes \begin{equation} \left\langle w^{\alpha}(T)\right\rangle^{1/\alpha}=L^{-p}D \end{equation} On the other hand, from what we stated in theorem 5 we know that \begin{equation} T=\frac{1} {L^{2p}\prod^{p}_{i=1}(1+\gamma^{\ast}_{1}\lambda^{(i)})}\;\; \mbox{ ,where} \end{equation} $\gamma^*_1$=$\gamma_1/L^2$ and by $T$ we mean $T^{(p)}$. Here we have included up to $O(\lambda)$ contributions to renormalized local times for scaling the running time of the process. In this scaling factor of the renormalization transformation, the $O(\lambda)$ contribution comes from the first term in right-hand side of eq(24). See Figure 4.\ {}From eq(54) and eq(53) follows \begin{equation} \left\langle w^{\alpha}(T)\right\rangle^{1/\alpha}=DT^{1/2} \left(\prod^{p}_{i=1}(1+\gamma^{\ast}_{1}\lambda^{(i)})\right)^{1/2} \end{equation} or \begin{equation} \left\langle w^{\alpha}(T)\right\rangle^{1/\alpha} \sim DT^{1/2}\left(e^{\gamma^{\ast}_{1}\sum^{(p)}_{i=1} \lambda^{(i)}}\right)^{1/2}. \end{equation} For $\beta=2$, $d=4$ and up to order $(\lambda^{(i)})^2$, follows \begin{equation} \lambda^{(i+1)}=\lambda^{(i)}-\beta_{1}(\lambda^{(i)})^{2}. \end{equation} Introducing the solution of the eq(57) recursion into eq(56) this becomes \begin{equation} \left\langle w^{\alpha}(T)\right\rangle^{1/\alpha} \sim DT^{1/2} e^{\frac{\gamma^{\ast}_{1}}{2\beta_{1}}ln p} \end{equation} or \begin{equation} \left\langle w^{\alpha}(T)\right\rangle^{1/\alpha} \sim DT^{1/2} (p)^{\frac{\gamma^{\ast}_{1}}{2\beta_{1}}} \end{equation} In eq(59) we have assumed p to be large enough so $\lambda^{-1}<<\beta_{1}(p)$. Taking the asymptotic limit we rewrite eq(59) as \begin{equation} \left\langle w^{\alpha}(T)\right\rangle^{1/\alpha} \sim DT^{1/2} log^{\frac{\gamma^{\ast}_{1}}{2\beta_{1}}}T, \end{equation} which is the asymptotic behavior of the end-to-end distance.\ It only remains to know the value of $\left(\frac{\gamma^{\ast}_{1}}{\beta_{1}}\right)$. Actually we can calculate $\gamma^{\ast}_{1}$ and $\beta_{1}$ from their definitions.\ Let us start with $\gamma_{1}$, from eq(37) we obtain \begin{equation} \gamma_{1}=\left\{ \left( \prod_{i'_{\alpha_{1}}\in J_{X'_{i'}}} e^{q_{1}(L^d-1)r(L^{\beta}t'_{i'_{\alpha_{1}}})} \right)^{-1}\times \right. \end{equation} $$ \times \left( \sum_{n_{i'}}\;L^{d}\int\prod_{i\in I_{X'_{i'}}} \;dt_{i}\;\prod_{i'_{\alpha_{1}}\in J_{X'_{i'}}} \;\delta (\sum^{m_{i'_{\alpha_{1}}}}_{{i}=m_{i'_{\alpha_{1}}-1}+1}\;\;t_{i}- L^{\beta }t'_{i'_{\alpha_{1}}})\times \right. $$ $$ \times \left( \begin{array}{c} n_{i'}+1 \\ 2 \end{array} \right) (q_{1})^{\left(n_{i'}-1\right)}r^{\left(n_{i'}\right)} (L^{d}-1)\times...\times (L^{d}-(n_{i'}-1)) \times $$ $$ \left. \times \sum_{X_{i}\in ({\it G}_{1})_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}\right\}\in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}}))} \chi\left(\left(i_{\alpha_{1}},j_{\alpha_{1}}\right)\in i'_{\alpha_{1}} \right) \right\}^{l.c.} $$ Hereafter, we assume $L^{d}>>n_{i'}$, so $$ \left(L^{d}-(n_{i'}-1)\right)\sim \left(L^{d}-1\right), $$ i.e. the number of points inside each ${\it G}_{1}$ coset is larger than the number of steps the walk $w'$ spends inside each L-block. Thus, the numerator of eq(61) can be written as \begin{equation} \frac{\left(\sum_{i'_{\alpha_{1}}\in J_{X'_{i'}}} L^{\beta}t' _{i'_{\alpha_{1}}}\right)^{2} } {2q_{1}(L^{d}-1)} \prod_{i'_{\alpha_{1}}\in J_{X'_{i'}}} \int^{1}_{0}\sum_{n_{i'}}\left(n_{i'}+1\right) \left(n_{i'}\right)^{2}\times \end{equation} $$ \times \left(n_{i'}-1\right) \frac{ \left( rq_{1}(L^{d}-1)(L^{\beta}t'_{i'_{\alpha_{1} } }) (1-t-t^{\ast})\right)^{n_{i'}}} {n_{i' }! } tdtt^{\ast}dt^{\ast}. $$ where we have taken $(1-t-t^{\ast})^{n_{i'}-2}$ $\sim $ $(1-t-t^{\ast})^{n_{i'}}$. To obtain an asymptotic estimate of eq(61), we assume that the following holds $$ \left(n_{i'}+1\right)\left(n_{i'}\right)^{2} \left(n_{i'}-1\right)\sim \frac{\left(n_{i'}\right)!} {\left(n_{i'}-4\right)!} $$ Althought from this follows $n_{i'}$ chosen to be large, we certainly assume finite local times after the renormalization-group transformation is applied.\ Substituting eq(62) in eq(61), with a jumping rate $r$ such that\newline $rq_{1}(L^{d}-1)\sim 1$ (as done in reference[5]) and for $\beta =2$ we obtain, in the asymptotic limit, $\gamma^{\ast}_{1}\sim 8$ provided $rq_{1}(L^{d}-1)\sim 1$.\ To calculate $\beta_{1}$ we use eq(40). \begin{equation} \beta_{1}=\frac{1}{2} \left\{ \left( \prod_{i'_{\alpha_{1}}\in J_{X'_{i'}}} e^{q_{1}(L^d-1)r(L^{\beta}t'_{i'_{\alpha_{1}}})} \right)^{-1}\times \right. \end{equation} $$ \times \left(L^{d} \sum_{(n_{i'_{a}})}\; \sum_{(n_{i'_{b}})}\; \int \prod_{i_{a}\in I_{X'_{i'}}} \;dt_{i_{a}}\; \prod_{i_{b}\in I_{X'_{i'}}} \;dt_{i_{b}}\; \right. \prod_{i'_{\alpha_{1}}\in J_{X'_{i'}}} \;\delta (\sum^{m_{i'_{\alpha_{1}}}}_{{i_{a}}=m_{i'_{\alpha_{1}}-1}+1}\;\; t_{i_{a}}- L^{\beta }t'_{i'_{\alpha_{1}}})\times $$ $$ \times \prod_{j'_{\alpha_{1}}\in J_{X'_{i'}}} \;\delta (\sum^{m_{j'_{\alpha_{1}}}}_{{i_{b}}=m_{j'_{\alpha_{1}}-1}+1}\;\; t_{i_{b}}- L^{\beta }t'_{j'_{\alpha_{1}}}) \left( \begin{array}{c} n_{i'_{a}}+1 \\ 2 \end{array} \right) \left( \begin{array}{c} n_{i'_{b}}+1 \\ 2 \end{array} \right) (q_{1})^{\left(n_{i'_{a}}\right)}r^{\left(n_{i'_{a}}\right)}\times $$ $$ \times (q_{1})^{\left(n_{i'_{b}}-1\right)}r^{\left(n_{i'_{b}}\right)} (L^{d}-1)\times...\times (L^{d}-(n_{i'_{a}}+n_{i'_{b}}-1)) \times $$ $$ \times \sum_{X_{i}\in ({\it G}_{1})_{X'_{i'}}} \sum_ {\stackrel{i_{\alpha_{1}}< j_{\alpha_{1}}< i_{\alpha_{2}}< j_{\alpha_{2}}} {\left\{i_{\alpha_{1}},j_{\alpha_{1}}, i_{\alpha_{2}},j_{\alpha_{2}}\right\} \in J_{X_{i}}}} t_{i_{\alpha_{1}}}t_{j_{\alpha_{1}}} {\bf 1}_{(w(t_{i_{\alpha_{1}}})=w(t_{j_{\alpha_{1}}})| i_{\alpha_{1}},j_{\alpha_{1}}\leq n_{i'_{a}})} \times $$ $$ \times \left. \left. t_{i_{\alpha_{2}}}t_{j_{\alpha_{2}}} {\bf 1}_{(w(t_{i_{\alpha_{2}}})=w(t_{j_{\alpha_{2}}})| i_{\alpha_{1}},j_{\alpha_{1}}\leq n_{i'_{b}})} \chi\left(\left(i_{\alpha_{1}},j_{\alpha_{1}}, i_{\alpha_{2}},j_{\alpha_{2}}\right)\in \left(i'_{\alpha_{1}}< j'_{\alpha_{1}}\right) \right) \right) \right\}^{l.c.} $$ As we did for the calculation of $\gamma_{1}$, we assume $L^{d}>>n_{i'}$, so \begin{equation} \left(L^{d}-(n_{i'_{a}}-1)\right)\sim (L^{d}-1)\;\mbox{ and }\; \left(L^{d}-(n_{i'_{b}})\right)\sim (L^{d}-1). \end{equation} Furthermore, we choose $ n_{i'_{b}}\sim n_{i'_{b}}-1, $ and approximations in eq(62) to hold for both $n_{i'_{a}}$ and $n_{i'_{b}}$ with a jumping rate $r$ such that $rq_{1}(L^{d}-1)\sim 1$. For $\beta=2$ in the asymptotic limit, we obtain $\beta_{1}\sim 32$ provided $rq_{1}(L^{d}-1)\sim 1$.\ Finally eq(60) becomes \begin{equation} \left\langle w^{\alpha}(T)\right\rangle^{1/\alpha} \sim (DT)^{1/2}log^{1/8}T \end{equation} Q.E.D.\ Note that $d=4$ is the only choice that renders eq(25) (for $\beta=2$) to a recursion as simple as eq(57) provided that $rq_{1}(L^{d}-1)\sim 1$.\ We want to remark that the heuristic study of the asymptotic end-to-end distance of a weakly SARW on a hierarchical lattice, $d=4$, is independent of hypotesis a) in the summary of former section. This is because we could have obtained the $O(\lambda)$ contribution to renormalized local times without introducing initial mass into the process.\ \section{Summary} In this paper we present a real space renormalization-group map, on the space of probabilities, to study weakly SARW that penalizes the (self-)intersection of two random walks for a hierarchical lattice, in dimension four. This hierarchical lattice has been labeled by elements of a countable, abelian group ${\it G}$. For any random function $F(w)$ on ${\it G}$ of the form described in Section 3, i.e. factorizable on the lattice, ( see eq(17) and eq(21) for examples of suitable $P(w)$) we can descend from the study of the space of walks on the whole lattice to the trajectory in the contracting ${\it G}_{1}$ cosets. Then we show how the L\'{e}vy process studied in reference [5] is a particular case of the processes that are (or flow to) fixed points of the renormalization-group map. We apply the renormalization-group map on some random walk models with configurational measure, working out explicitly the weakly SARW case. An heuristic proof of the end-to-end distance for a weakly SARW on a hierarchical lattice is derived. This gives a new probabilistic meaning to the exponent of the logarithmic correction.\ In Section 4 we study a weakly SARW that penalizes the (self-)intersection of two random walks. The weakly SARW probability studied, involves a factor linear in local times, i.e. a random walk representation of a field-theoretical gaussian component that adds to the corresponding term produced for the renormalization-group map applied on the weakly SARW. We show how this probability flows to a fixed form (the random walk version of the field-theoretical result given in reference [5]) relying on;\ a) An hypothesis that assumes we can approach the expectation of the interaction energy in terms of the product of expectations for each of its factors, conditioned to applying the renormalization-group map.\ b) The hierarchical metric space used to label the lattice that allows, for $F(w)$ of the form described in Section 3 (f.e. eq(17) and eq(21)), the factorization in terms of the quotient group ${\it G}/{\it G}_{1}$ and the image of the renormalization-group map on the cosets ${\it G}_{1}$, each step the renormalization-group is applied.\ c) A class of realizations of the model such that, each step we apply the renormalization-group transformation, the renormalized fixed walk visits $1, 2, 3$ times different sites in the lattice ${\it G}/{\it G}_{1}$, at least once. Other realizations might not allow us to study the flow of the (self-)intersecting coefficients that are interesting for us.\ d) A formal Taylor series expansion in $\lambda$ from which, upon renormalization, we use only the linear contributions. \ Our result improves the field-theoretical approach \cite{Ev} by obtaining an exact expression for the parameters that appear in the flow of $\lambda$ and $\xi$. This is a crucial feature used to obtain heuristically the asymptotic behavior of the end-to-end distance for a weakly SARW that penalizes the (self-)intersection of two random walks. Furthermore, the method here presented is full of physical intuition and suitable of being applied to discrete kinetically growing measure models.\ Following standard thinking we shrink all space and time magnitudes each step the map is applied. We shrink time taking into account $O(\lambda)$ contributions to renormalized local times, generated by applying the renormalization-group map to the weakly SARW that penalizes (self-) intersections of walks. Length type magnitudes are shrunk as usual. We present this, as a possible origin for the expression $\left\langle w^{\alpha}(T)\right\rangle^{1/\alpha}\sim (DT)^{1/2}log^{1/8}T$ as T tends to infinity, in $d=4$, for a weakly SARW that penalizes the (self-)intersection of two random walks on a hierarchical lattice.\ \vspace*{10mm} \section{Acknowledgments} This work was partially supported by CONACYT REF 4336E9406, M\'{e}xico. I wish to thank S.N. Evans, Department of Statistics, UC Berkeley, for helpful discussion and ITD, UC Davis for its hospitality. \baselineskip0.6cm
\section{Introduction} \label{sec:Introduction} The much heralded convergence of the Standard Model gauge couplings in supersymmetric Grand Unified Theories (GUTs) \cite{EKN}, is continually being challenged by ever more precise LEP measurements of the gauge couplings. It was realized early on that the effect of the GUT particles responsible for the onset of the unified theory, was not negligible \cite{GUTthresholds}. However, because of the presumed great uncertainty in the GUT physics, such discussions have been largely carried out in the context of the minimal $SU(5)$ supergravity model \cite{CAN}. Central to the study of these issues is the technical point of how exactly these GUT (or lighter) particles decouple at scales below their masses. Recent investigations \cite{smooth} reveal that a ``smooth" decoupling of particles leads to noticeable differences from the step-function approximation. Moreover, these new effects coupled with the latest LEP data on $\sin^2\theta_W$ and the determination of the top-quark mass, lead to a greatly increased prediction for $\alpha_3(M_Z)$ \cite{Clavelli,Bagger,LP}, strongly suggesting that minimal $SU(5)$ GUT thresholds are unable to bring the $\alpha_3(M_Z)$ prediction down to the experimentally acceptable range \cite{Clavelli}. This impasse may be resolved with a significant contribution from Planck-scale non-renormalizable operators \cite{NROs,LP}, although such effects call into question the whole field-theoretical approximation to the gauge coupling unification problem. Even if Planck-scale physics can resolve the present $\alpha_3$ discrepancy in minimal $SU(5)$, this GUT model suffers from a well known fine-tuning \cite{DG} regarding the solution of the doublet-triplet splitting problem of the Higgs pentaplets. At least three solutions to this problem (all involving non-minimal GUT models) have been proposed: the missing-partner mechanism \cite{MPM}, the sliding-singlet mechanism \cite{SSM}, and the pseudo-goldstone-boson mechanism \cite{PGBM}. In the sliding-singlet mechanism radiative corrections destroy the gauge hierarchy \cite{Lahanas}, whereas an additional global or local $SU(6)$ symmetry is required in the pseudo-goldstone-boson mechanism. It is very suggestive that the same investigations that uncover the $\alpha_3$ discrepancy in minimal $SU(5)$, also show that in the so-called Missing Doublet Model (MDM) \cite{MPM}, which has as its central component the missing-partner mechanism, the $\alpha_3$ prediction is decreased to acceptable values \cite{Bagger,Clavelli}. As we discuss below, some variants of the missing-doublet model \cite{MNTY2,HMTY} also solve the problematic situation with dimension-five proton decay operators in minimal $SU(5)$, which require the Higgs triplet mass ($M_{H_3}$) to exceed the GUT scale and the supersymmetric spectrum to be tuned in specific ways \cite{AN,HMY}, especially when cosmological constraints are simultaneously enforced \cite{LNP}. In fact, an updated analysis has recently shown \cite{HMTYpd} that the upper bound on the Higgs triplet mass from unification constraints ({\em i.e.}, $M_{H_3}\propto e^{-5\pi/3\alpha_3},\ \alpha_3<\alpha_3^{\rm max}$), and the corresponding lower bound from proton decay constraints ({\em i.e.}, $\tau_p\propto M^2_{H_3},\ \tau_p>\tau^{\rm min}_p$) are very close to each other, leaving only a small window of allowed parameter space in minimal $SU(5)$. Note also that, because of the rather large representations introduced in the MDM (\r{75},\r{50},\rb{50}), it is necessary to assign some of these Planck-size masses, in order to avoid the onset of a strongly-interacting GUT below the Planck scale \cite{Clavelli,HMTY}. Thus, Planck-scale physics is again unavoidable in this more realistic version of $SU(5)$ GUTs. $SO(10)$ GUTs \cite{GFM,GN} have also received a great deal of attention lately \cite{BB,Mohapatra,Raby,inspired}, with interesting successes in the area of quark and lepton masses and mixings, although the $\tan\beta={\cal O}(50)$ prediction requires fine-tuning of the supersymmetric spectrum \cite{NR,Carena} to reconcile it with radiative electroweak breaking. Assuming universal soft supersymmetry breaking, the further constraints from $B(b\rightarrow s\gamma)$ and cosmology strongly disfavor the model \cite{BOP}. However, most of these shortcomings are overcome when one allows certain classes of non-universal scalar masses \cite{OP,BOP}. More to the point, the successes of $SO(10)$ rely on the existence of certain non-renormalizable operators (as originally suggested in Ref.~\cite{EG+NS}) that are presumed to be obtained from a string-derived model at the Planck scale. Despite initial claims \cite{Lykken}, no {\em consistent} $SO(10)$ GUT string model has been derived in the context of free-fermionic strings \cite{Cleaver}. However, these failed attempts have been enough to fuel a series of ``string-inspired" $SO(10)$ GUT models \cite{inspired}, which are limited to certain type and number of representations (those allowed by the level-two Kac-Moody construction\footnote{At level two, the allowed unitary massless representations are \r{1},\r{10},\r{16},\rb{16},\r{45},\r{54} \cite{ELN}. A string model containing the \r{126},\rb{126} representations used in traditional $SO(10)$ model building requires an unlikely level-five construction \cite{ELN}.}), forcing model builders to rely heavily on postulated effective non-renormalizable operators \cite{inspired}. Level-two string $SO(10)$ GUT models have been consistently constructed in the context of symmetric orbifolds \cite{Aldazabal}, but with limited phenomenological success, especially in dealing with the doublet-triplet splitting problem. In view of its field-theoretical successes, in this paper we revisit the missing-doublet model as a well-motivated, realistic contender for a grand unified model. We first review the original MDM and its features and shortcomings (Sec.~\ref{sec:observable}). We then propose a simple extension of the model to naturally suppress dimension-five proton decay operators (Sec.~\ref{sec:observable}). Our most substantive contribution is to endow this {\em supergravity} model with a hidden sector containing gauge and matter degrees of freedom (Sec.~\ref{sec:hidden}). Hidden sector gaugino condensation triggers supersymmetry breaking which, as we discuss, can be of the desired magnitude for suitable choices of the hidden gauge group and hidden matter spectrum. The matter condensates provide a new dynamical intermediate scale which, via non-renormalizable interactions, generates a low-energy Higgs mixing term $\mu$. With the introduction of right-handed neutrinos to the model, this scale also becomes their mass scale, which provides a suitable see-saw spectrum of neutrino masses (Secs.~\ref{sec:observable},\ref{sec:CBA}). We show that the model is consistent with gauge coupling unification for experimentally acceptable values of $\alpha_3(M_Z)$ and that dimension-five proton decay operators are consistent with present limits even for large values of $\tan\beta$ (Sec.~\ref{sec:unification}). Also, the out-of-equilibrium decays of the right-handed neutrinos subsequent to inflation produce a lepton asymmetry which is re-processed into a baryon asymmetry by strongly-interacting Standard Model effects (sphalerons) at the electroweak scale (Sec.~\ref{sec:CBA}). We finally compare the features of this traditional GUT model with that of the readily string-derivable $SU(5)\times U(1)$ model, and discuss the prospects of deriving the revamped MDM from string theory (Sec.~\ref{sec:comparison}). We summarize our conclusions in Sec.~\ref{sec:conclusions}. \section{The observable sector} \label{sec:observable} The original MDM \cite{MPM} can be described by the following set of observable sector fields: $\Sigma$ (\r{75}), $\theta$ (\r{50}), $\bar\theta$ (\rb{50}), $h$ (\r{5}), $\bar h$ (\rb{5}), $F_{1,2,3}$ (\r{10}'s), $\bar f_{1,2,3}$ (\rb{5}'s), interacting via the superpotential \begin{equation} W=\coeff{1}{2}M_{75}\,\Tr{\Sigma^2}+\coeff{1}{3}\lambda_{75}\,\Tr{\Sigma^3} +\lambda_4\,\bar\theta\Sigma h+\lambda_5\,\theta\Sigma\bar h +M_{50}\,\theta\bar\theta+\lambda^{ij}_2 F_iF_jh +\lambda^{ij}_1 F_i\bar f_j \bar h\ . \label{eq:W} \end{equation} The expectation value of the \r{75} can be chosen such that the $SU(5)$ gauge symmetry is broken down to the Standard Model one, in which case the scalar potential that follows from Eq.~(\ref{eq:W}) gives $\vev{\Sigma}\sim M_{75}/\lambda_{75}$. The $\bar\theta\Sigma h$, $\theta\Sigma\bar h$, and $\theta\bar\theta$ terms in $W$ effect the doublet-triplet mechanism via the mass matrix \begin{equation} \bordermatrix{ &\bar h_3&\bar\theta_3\cr h_3&0&\lambda_4\vev{\Sigma}\cr \theta_3&\lambda_5\vev{\Sigma}&M_{50}}\ , \label{eq:2/3} \end{equation} where the subscript `3' indicates the $SU(2)_L$ singlet, $SU(3)_C$ triplet component of the corresponding $SU(5)$ representation. This matrix clearly yields massive ($\sim\vev{\Sigma}\sim M_{\rm GUT}$) Higgs triplets ($h_3,\bar h_3$), whereas the doublets ($h_2,\bar h_2$) remain massless. The $M_{50}$ term is not required for a successful doublet-triplet splitting. However, it is introduced in order to give large masses to the many leftover components of the \r{50},\rb{50} representations. The last two terms in W (\ref{eq:W}) provide the Yukawa matrices for the Standard Model fermions, implying the usual relations ({\em e.g.}, $\lambda_b=\lambda_\tau$). Despite the above natural solution to the doublet-triplet splitting problem, the magnitude of the dimension-five ($d=5$) proton decay operators still needs to be assessed. The crucial element in this calculation is the effective $h_3\bar h_3$ mixing term. If in Eq.~(\ref{eq:W}) $M_{50}$ were allowed to vanish, then there would be no mixing whatever, and the $d=5$ operators would be negligible. In practice $M_{50}$ cannot vanish, and we are left with two possibilities: (i) $M_{50}\sim \vev{\Sigma}$, and (ii) $M_{50}\gg\vev{\Sigma}$. The first case implies an effective Higgs-triplet mixing term of the same magnitude as in the minimal $SU(5)$ model, and therefore similar difficulties in suppressing proton decay. However, this case is not really viable since above the GUT scale the large \r{50},\rb{50} representations increase the $SU(5)$ beta function so much that the gauge coupling becomes non-perturbative before reaching the Planck scale \cite{Clavelli,HMTY}. We are left with the second alternative with $M_{50}\sim M$, where $M=M_{Pl}/\sqrt{8\pi}\approx10^{18}\,{\rm GeV}$ is the appropriate gravitational scale. Unfortunately, this choice leads to a see-saw type mass for the Higgs triplets: $m_{h_3,\bar h_3}\sim \vev{\Sigma}^2/M\sim10^{14}\,{\rm GeV}$, and effective mixing of the same magnitude, which makes proton decay much too fast. Various variants of the MDM have been proposed to deal with the proton decay problem in a more effective way \cite{MPM,MNTY2,HMTY}. Here we follow the suggestion in Ref.~\cite{HMTY}, whereby the following additional fields are introduced: $\theta'$ (\r{50}), $\bar\theta'$ (\rb{50}), $h'$ (\r{5}), $\bar h'$ (\rb{5}). The superpotential for the model is that in Eq.~(\ref{eq:W}) with $M_{50}\equiv0$, and supplemented by \begin{equation} W'=\lambda_4'\,\bar\theta'\Sigma h'+\lambda_5'\,\theta'\Sigma\bar h' +M'_{50}\,\theta\bar\theta'+M'_{50}\,\theta'\bar\theta\ , \label{eq:W'} \end{equation} where we again take $M'_{50}\sim M$. These interactions lead to the following generalized Higgs-triplet mass matrix \begin{equation} \bordermatrix{ &\bar h'_3&\bar\theta_3&\bar h_3&\bar\theta'_3\cr h_3&0&\lambda_4\vev{\Sigma}&0&0\cr \theta'_3&\lambda'_5\vev{\Sigma}&M'_{50}&0&0\cr h'_3&0&0&0&\lambda'_4\vev{\Sigma}\cr \theta_3&0&0&\lambda_5\vev{\Sigma}&M'_{50}}\ , \label{eq:2/3'} \end{equation} and effective interactions \cite{HMTY} \begin{equation} \left(\lambda_4\lambda'_5{\vev{\Sigma}^2\over M'_{50}}\right)h_3\bar h'_3+ \left(\lambda'_4\lambda_5{\vev{\Sigma}^2\over M'_{50}}\right)h'_3\bar h_3 \equiv M_{H_3}\,h_3\bar h'_3+M_{\bar H_3}\,h'_3\bar h_3\ . \label{eq:mixings} \end{equation} Since there is no effective interaction between $h_3$ and $\bar h_3$ (the only triplets that interact with the Standard Model fermions), the $d=5$ operator is negligible. If the superpotential $W+W'$ were the complete model, we would have managed to make all the non-minimal fields sufficiently heavy or non-interacting. However, we would have left two pairs of Higgs doublets $h_2,\bar h_2$ and $h'_2,\bar h_2'$ with no apparent use for the second pair, and if light, with severe trouble with gauge coupling unification. Let us assume the existence of a mass term $M_{h'}h'\bar h'$, with no specific origin for $M_{h'}$ for now. Such a term contains $M_{h'}h'_3\bar h'_3$, which ``hooks up" the two disconnected pieces in Eq.~(\ref{eq:mixings}) and allows $d=5$ proton decay to occur, with an operator proportional to \begin{equation} {1\over M_{H_{\rm eff}}}\equiv{M_{h'}\over M_{H_3} M_{\bar H_3}} \sim {M_{h'}\over [\vev{\Sigma}^2/M'_{50}]^2}\ , \label{eq:Heff} \end{equation} where $M_{H_{\rm eff}}$ is the effective Higgs triplet mass. Since in the minimal $SU(5)$ model with $M_{H_{\rm eff}}=M_{H_3}\mathrel{\mathpalette\@versim>} 10^{17}\,{\rm GeV}$, the present experimental bounds on proton decay are satisfied without strong restrictions on the parameter space \cite{AN,HMY}, we effectively require $M_{h'}\lsim10^{11}\,{\rm GeV}$. But where does this intermediate scale come from? It has been suggested that this scale could be generated dynamically via the breaking of a Peccei-Quinn symmetry \cite{MSY,HMTY}. A more modern and economical approach to the generation of intermediate scales, especially in the context of supergravity, is to consider the condensation of a hidden sector gauge group that triggers supersymmetry breaking. Non-renormalizable interactions coupling hidden sector matter fields to observable fields may then naturally generate the intermediate scale.\footnote{This mechanism is commonly available in string model building \cite{decisive}.} Specifically, we add to our model the following superpotential terms\footnote{The apparent asymmetry between the $h\bar h$ and $h'\bar h'$ couplings may be understood on the basis of additional local $U(1)$ quantum numbers, which are broken near the Planck scale and are carried by both hidden and observable sector particles, as is common in string model building \cite{decisive}. For further symmetry arguments motivating these choices, see {\em e.g.}, Ref.~\cite{CKN}.} \begin{equation} W''= \lambda_7\, h\bar h\,{(T\bar T)^2\over M^3} +\lambda'_7\, h'\bar h'\,{T\bar T\over M}\ , \label{eq:W''} \end{equation} where $T\bar T$ is a gauge-singlet hidden-sector composite ({\em e.g.}, \r{4}\rb{4} in $SU(4)$). When the hidden sector condenses, we generate dynamically two mass scales: \begin{equation} M_{h'}=\lambda'_7\,{\vev{T\bar T}\over M}\,,\qquad \mu=\lambda_7 \,{\vev{T\bar T}^2\over M^3}\ . \label{eq:scales} \end{equation} Note that for $\vev{T\bar T}/M\sim10^{10}\,{\rm GeV}$, we would obtain for the masses of the extra pair of doublets $M_{h'}\sim10^{10}\,{\rm GeV}$, and an effective Higgs-triplet mixing which satisfies proton decay constraints automatically. We would also obtain dynamically\footnote{This dynamical generation of the $\mu$ parameter via non-renormalizable interactions is also familiar from string model-building \cite{decisive,Casasmu,CKN}.} a very desirable Higgs mixing parameter $\mu\sim100\,{\rm GeV}$. In the next section we explore the hidden sector of the model with these phenomenological constraints in mind. One of the main model-building shortcomings of $SU(5)$ GUTs is the not-so-obvious source of neutrino masses. In fact, neutrino masses can be introduced by simply adding right-handed neutrino ($SU(5)$ singlet) fields to the model. To implement the standard see-saw mechanism we introduce three singlet fields $\nu^c_{1,2,3}$ with the following superpotential \begin{equation} W'''=\lambda^{ij}_3\, \bar f_i \nu^c_j h + \lambda^{ij}_6\, \nu^c_i\nu^c_j \, {T\bar T\over M}\ . \label{eq:W'''} \end{equation} After hidden sector condensation and electroweak symmetry breaking, we obtain the following see-saw neutrino mass matrix \begin{equation} \bordermatrix{&\nu_j&\nu^c_j\cr \nu_i&0&\lambda_3^{ij}v_2\cr \nu^c_i&\lambda_3^{ji}v_2&\lambda^{ij}_6\vev{T\bar T}/M}\ , \label{eq:see-saw} \end{equation} and light neutrino see-saw masses $m_{\nu}\sim\lambda^3_2 v^2_2/[\lambda_6\vev{T\bar T}/M]$. For simplicity, in what follows we assume $\lambda^{ij}_6=\lambda^i_6\delta_{ij}$. With our above desired value of $\vev{T\bar T}/M\sim M_{\nu^c}\sim10^{10}\,{\rm GeV}$, and $\lambda_3 v_2\sim10\,{\rm GeV}$, typical see-saw light neutrino masses follow, {\em i.e.}, $m_{\nu_\tau}\sim10\,{\rm eV}$. Further discussion of the consequences of this see-saw matrix for the light neutrino masses and mixing angles is given in Sec.~\ref{sec:CBA} below. \section{The hidden sector} \label{sec:hidden} Our supergravity model is endowed with a hidden sector which communicates with the observable sector via gravitational interactions (or via $U(1)$ gauge interactions broken near the Planck scale). The hidden sector consists of a hidden gauge group and a set of matter representations, which for convenience we take to be $SU(N_c)$ with $N_f$ flavors ($T_i,\bar T_i,\ i=1\rightarrow N_f$) and $N_f<N_c$. This gauge group starts with a gauge coupling $g$ at the Planck scale ($Q=M$), and becomes strongly interacting at the condensation scale defined by \begin{equation} \Lambda=M e^{8\pi^2/\beta g^2}\ , \label{eq:Lambda} \end{equation} where the beta function is given by $\beta=-3N_c+N_f$. For simplicity we assume that all the flavors are ``light", {\em i.e.}, they have masses\footnote{Massless flavors lead to pathologies ({\em i.e.}, no vacuum), which can nonetheless be remedied by invoking supersymmetry-breaking masses for the $T_i,\bar T_i$ fields \cite{Peskin}.} $m\ll\Lambda$. At the condensation scale, the strongly interacting theory is described in terms of composite ``meson" fields $T_i\bar T_i$. The dynamics of this system can be obtained from an effective Lagrangian with the following non-perturbative superpotential \cite{Seiberg} \begin{equation} W_{\rm non-pert}=(N_c-N_f)\,{\Lambda^{(3N_c-N_f)/(N_c-N_f)}\over (\det T\bar T)^{1/(N_c-N_f)}}+\Tr(mT\bar T)\ . \label{eq:Wnp} \end{equation} Minimization of the corresponding scalar potential results in the following expectation values for the mesons fields $\vev{T\bar T}$ (we work in a diagonal flavor basis) \begin{equation} \vev{T\bar T}=\Lambda^{(3N_c-N_f)/N_c}\,(\det m)^{1/N_c}\, m^{-1} =\Lambda^3 \left({m\over\Lambda}\right)^{N_f/N_c}{1\over m} =\Lambda^2\, x^{(N_f/N_c)-1}\ , \label{eq:TT} \end{equation} where in the last expression we have defined $x\equiv m/\Lambda$, with $x<1$. Inserting this expectation value in $W_{\rm non-pert}$, we obtain \begin{equation} \vev{W}=N_c\, \Lambda^3\, x^{N_f/N_c}\ , \label{eq:Wvev} \end{equation} where $W$ includes all perturbative and non-perturbative contributions. In a supergravity theory, the scale of supersymmetry breaking is determined by the gravitino mass: $m_{3/2}=\vev{e^K\,W}$, where $K$ is the K\"ahler potential. In simple models $K=\sum\phi_i\phi_i^\dagger$, and thus $\vev{K}=0$. More complicated forms of $K$ are obtained in string models (where the dilaton and moduli fields play an important role). For our present purposes, we simply assume that $\vev{e^K}\sim1$. This implies that $\vev{W}$ is the sole source of supersymmetry breaking, {\em i.e.}, \begin{equation} m_{3/2}\sim \vev{W}\sim \left({\Lambda\over M}\right)^3\, x^{N_f/N_c}\, M\ , \label{eq:m3/2} \end{equation} where we have restored the units in the expression for $m_{3/2}$. With the results in Eqs.~(\ref{eq:TT}) and (\ref{eq:m3/2}) for $\vev{T\bar T}$ and $m_{3/2}$, we can now investigate the conditions on $N_c$, $N_f$, and $x$ that would yield the desired results: $\vev{T\bar T}/M=10^p\,{\rm GeV}$ and $m_{3/2}=10^q\,{\rm GeV}$, with $p\sim10$ and $q\sim3$. In terms of $p$ and $q$, we can solve simultaneously Eqs.~(\ref{eq:TT}), (\ref{eq:m3/2}), and (\ref{eq:Lambda}), to obtain \begin{equation} N_c={p-q\over18-q}\,N_f+{8\pi^2\over g^2}\,\,{\log_{10} e\over18-q}\ , \label{eq:Nc} \end{equation} and \begin{equation} x=10^{2N_c\,({3\over2}p-q-9)/\beta}\ . \label{eq:x} \end{equation} Thus, for a given value of $g$ and $N_f$, we obtain $N_c$ (and thus $\beta$) from Eq.~(\ref{eq:Nc}). With this value of $N_c$, $x$ is determined from Eq.~(\ref{eq:x}), and $\Lambda$ from Eq.~(\ref{eq:Lambda}). For the desired $p=10$ and $q=3$, and with the sensible inputs $N_f=1$ and $g=0.7$, we obtain $N_c=5$, $x\approx0.01$, and $\Lambda\approx10^{13}\,{\rm GeV}$. That is, an $SU(5)$ hidden gauge group with one light flavor. The general constraints on $N_c$ and $N_f$ for given values of $g$ are shown in Fig.~\ref{fig:NcNf}, for $q=2\to3$ ({\em i.e.}, $m_{3/2}=100\,{\rm GeV}\to1\,{\rm TeV}$) and $p=10$. We do not address here the calculation of the observable-sector soft-supersymmetry-breaking scalar and gaugino masses, since these depend on the specific choices for the K\"ahler function and the gauge kinetic function, although their overall scale is already determined by $m_{3/2}$. The ``flat" choice $K=\sum\phi_i\phi^\dagger_j$ leads to the usual universal scalar masses, but this choice is not unique. \section{Unification and proton decay} \label{sec:unification} The revamped MDM presented in the two previous sections contains several departures from conventional gauge coupling unification: (i) there is a pair of Higgs doublets with intermediate-scale masses ($M_{h'}\sim10^{10}\,{\rm GeV}$), (ii) there is a richer structure of GUT particles, including two pairs of Higgs triplets (from the \r{5},\rb{5} representations) with masses $M_{H_3,\bar H_3}\sim10^{14}\,{\rm GeV}$, and (iii) there is a spectrum of masses for the different components of the \r{75}, all near the unification scale. There is also a hidden gauge group, with an in-principle independent gauge coupling at the Planck scale (denoted by $g$ in Sec.~\ref{sec:hidden}).\footnote{In the spirit of string unified models one could assume that the observable and hidden gauge couplings are related at the Planck scale or at the string scale ($M_{\rm str}\approx4\times10^{17}\,{\rm GeV}$).} Fortunately, the issue of gauge coupling unification in the observable sector has already been addressed in detail in Ref.~\cite{HMTY}. Those calculations are directly applicable to our model since the observable matter content and spectrum of the two models is the same, even though the dynamics providing the intermediate scale are different. Thus, here we limit ourselves to a brief summary of the relevant issues. Writing down the one-loop RGEs for the gauge couplings, including a common supersymmetric threshold at $M_{\rm SUSY}$, one can eliminate the unified $SU(5)$ coupling and obtain the following two relations \cite{HMTY,Method} \begin{eqnarray} \left(3\alpha^{-1}_2-2\alpha^{-1}_3-\alpha^{-1}_1\right)(M_Z)&=& {1\over2\pi}\left\{{12\over5}\ln{M_{H_3}M_{\bar H_3}\over M_{h'}M_Z} -2\ln{M_{\rm SUSY}\over M_Z}-23.3\right\} \label{eq:Relation1}\\ \left(5\alpha^{-1}_1-3\alpha^{-1}_2-2\alpha^{-1}_3\right)(M_Z)&=& {1\over2\pi}\left\{36\ln{(M^2_V M_\Sigma)^{1/3}\over M_Z} +8\ln{M_{\rm SUSY}\over M_Z}+12.1\right\} \label{eq:Relation2} \end{eqnarray} In these relations, $M_V=3\sqrt{15}(g_5/\lambda_{75})M_{75}$ is the mass of the GUT gauge bosons, and the explicit constants come from the splittings of the \r{75} relative to the $M_\Sigma=5M_{75}$ mass of its (\r{8},\r{3}) component. The above relations can be made more accurate by the inclusion of realistic low-energy supersymmetric thresholds, two-loop gauge coupling RGEs, and smooth decoupling of heavy particles. Once this is done, and the latest values of the Standard Model gauge couplings are input ({\em i.e.}, $\alpha^{-1}=127.9\pm0.2$, $\sin^2\theta_W=0.2314\pm0.0004$, $\alpha_3=0.118\pm0.007$), one obtains the following 1$\sigma$ allowed intervals \cite{HMTY}: \begin{eqnarray} 1.4\times10^{17}\,{\rm GeV}\le&{M_{H_3}M_{\bar H_3}\over M_{h'}}& \le 5.5\times10^{20}\,{\rm GeV}\ , \label{eq:H3range}\\ 8.4\times10^{15}\,{\rm GeV}\le&\left(M^2_V M_\Sigma\right)^{1/3}& \le2.6\times10^{16}\,{\rm GeV}\ . \label{eq:Vrange} \end{eqnarray} It is evident that our choices above, {\em i.e.}, $M_{H_3}\sim M_{\bar H_3}\sim 10^{14}\,{\rm GeV}$ and $M_{h'}\sim10^{10}\,{\rm GeV}$, are perfectly consistent with the constraint in Eq.~(\ref{eq:H3range}). The same is true for the middle-of-the-road choice $M_V\sim M_\Sigma$ ({\em i.e.}, $\lambda_{75}\sim g_5$), which yields a GUT scale close to $10^{16}\,{\rm GeV}$. We recall that we have set the masses of the \r{50},\rb{50} representations at the gravitational scale $M\approx10^{18}\,{\rm GeV}$ in order to prevent the onset of a non-perturbative $SU(5)$ regime below the Planck scale. Nonetheless, because of the needed GUT-scale \r{75} representation, the unified gauge coupling grows above the unification scale. However, it has been demonstrated that this coupling remains in the perturbative regime, {\em i.e.}, $\alpha\lsim0.1$ \cite{HMTY}. One could assume that the corresponding gauge coupling at the gravitational scale ($g\approx1$) is related to the gauge coupling from the hidden sector gauge group discussed in Sec.~\ref{sec:hidden}, as would be the case in string models. This relation would help to further constrain the viable hidden sector choices. For instance, assuming a ``super-unified" situation, where hidden and observable gauge couplings are equal near the gravitational scale, the constraints on the hidden sector choice can be read off Fig.~\ref{fig:NcNf} ($g=1$ curves). Concerning proton decay, gauge-boson-mediated dimension-six operators depend on $1/M^2_V$. From Eq.~(\ref{eq:Vrange}), $M_V$ is not expected to be much below $10^{16}\,{\rm GeV}$, unless $\lambda_{75}\gg g_5$, but this case is unlikely since $\lambda_{75}$ would be in the non-perturbative regime. Thus, we don't expect a particular enhancement of dimension-six operators in this model. More interesting is the situation with the dimension-five proton decay operators, which depend on the effective Higgs triplet mass ($M_{H_{\rm eff}}$) defined in Eq.~(\ref{eq:Heff}). The dominant proton partial lifetime is given by \cite{HMY,HMTYpd} \begin{equation} \tau(p\rightarrow K^+\bar\nu_\mu)=2.0\times10^{31}\,{\rm y} \left| {0.0056\,{\rm GeV}^3\over\beta}\,{0.67\over A_S}\,{\sin2\beta\over 1+y^{tK}}\, {M_{H_{\rm eff}}\over 10^{17}\,{\rm GeV}}\,{\,{\rm TeV}^{-1}\over f}\right|^2\ , \label{eq:pdecay} \end{equation} where $\beta=(5.6\pm0.8)\times10^{-3}\,{\rm GeV}^3$ is the relevant hadronic matrix element, $A_S$ is the short-distance renormalization factor, and $y^{tK}$ parametrizes the contribution of the third family relative to the first two ($|y^{tK}|\approx2$ for $m_t=175\,{\rm GeV}$) with an undetermined phase. The $f$ functions are the one-loop integrals which behave as $1/f\approx m^2_{\tilde q}/m_{\widetilde W}$ for $m_{\tilde q}\gg m_{\widetilde W}$. {}From unification constraints, Eq.~(\ref{eq:H3range}) indicates that $M_{H_{\rm eff}}>1.4\times 10^{17}\,{\rm GeV}\approx10M_V$. In this case, Eq.~(\ref{eq:pdecay}) and Ref.~\cite{AN} show that the present Kamiokande limit $\tau(p\rightarrow \bar\nu K^+)>1\times10^{32}\,{\rm y}$ \cite{PDG}, is satisfied provided $\tan\beta$ is not too large ($\tan\beta\lsim5$) and the universal scalar mass $m_0>300\,{\rm GeV}$. On the other hand, in our model we obtain $M_{H_{\rm eff}}\gsim10^{18}\,{\rm GeV} \approx100M_V$, and the experimental limit is satisfied rather comfortably, even for large values of $\tan\beta$ and presently accessible supersymmetric particle masses. For instance, for $m_{\tilde q}\approx 300\,(600)\,{\rm GeV}$ and $m_{\widetilde W}\approx 80\,{\rm GeV}$, $\tan\beta\lsim10\,(40)$ is required. Thus, $p\rightarrow \bar\nu K^+$ remains the dominant mode for proton decay, with good prospects for observation at the upcoming SuperKamiokande experiment and the proposed ICARUS facility. Note that the much-weakened proton-decay upper-bound on $\tan\beta$ offers a new possibility in the study of Yukawa coupling unification in $SU(5)$ GUTs ({\em i.e.}, $\lambda_b=\lambda_\tau$), which now also allow the so-called ``large-$\tan\beta$" solution \cite{YU}. \section{Cosmic baryon asymmetry} \label{sec:CBA} With the realization of significant electroweak baryon number violation at high temperatures, which occurs through ($B$+$L$)-violating but ($B$--$L$)-conserving non-perturbative sphaleron interactions \cite{sphalerons}, several new mechanisms for generating the cosmic baryon asymmetry have been proposed \cite{Olive}. These mechanisms produce a primordial lepton asymmetry (leptogenesis), which is then recycled by sphaleron interactions into a baryon asymmetry at the electroweak scale. It is important to note that primordial ($B$--$L$)-conserving asymmetries, such as those produced in traditional $SU(5)$ GUT baryogenesis, are likely to be wiped out by the sphaleron interactions \cite{erasure}. Therefore, in the context of $SU(5)$ GUTs, the leptogenesis-based mechanisms may be unavoidable. Here we consider the simplest of these mechanisms, based on the out-of-equilibrium decay of right-handed neutrinos, as first suggested in Ref.~\cite{FY},\footnote{Before the realization of the importance of the sphaleron interactions, Ref.~\cite{MNTY2} pointed out the possibility of generating a baryon asymmetry in the decay of right-handed neutrinos via baryon number violating GUT interactions.} and extended to supersymmetry in Refs.~\cite{CDO,MSYY}, and to $SU(5)\times U(1)$ GUTs in Refs.~\cite{ENO,ELNO}. We note that the lepton-asymmetric decays of right-handed sneutrino condensates \cite{AD,sneutrinos}, may provide an additional contribution to the lepton asymmetry that we discuss below. In order to satisfy the out-of-equilibrium condition in the decay of the right-handed neutrinos, one could follow the standard procedure of demanding that the $\nu^c_{1,2,3}$ decay rate be less than the expansion rate of the Universe at the time of $\nu^c$ decay. This condition leads to constraints on the $\lambda_3$ couplings of the right-handed neutrinos, that can be undesirable when trying to use the same couplings to compute the corresponding light see-saw neutrino masses. Even more problematic can be the need to obtain the surviving lepton asymmetry solely from the decays of the lightest right-handed neutrino ($\nu^c_1$), since the asymmetry produced in the decays of $\nu^c_{2,3}$ is typically wiped out by the $\nu^c_1$ interactions. Such potential difficulties have been exemplified in Ref.~\cite{ELNO}. Instead, here we follow an alternative scenario \cite{NOS},\footnote{Below we show that in our model, the traditional out-of-equilibrium scenario is also viable.} whereby the right-handed neutrinos are produced in the decays of the inflaton subsequent to inflation. The COBE data on the anisotropy of the cosmic microwave brackground radiation, interpreted in the context of inflation, allows one to deduce the inflaton mass to be $m_{\eta}\sim 10^{11}\,{\rm GeV}$ and the reheating temperature $T_R\sim10^8\,{\rm GeV}$ \cite{CDO}. The $\nu^c$ then decay immediately after inflation and out of equilibrium at the temperature $T_R\ll M_{\nu^c}$, {\em as long as} $M_{\nu_c}<m_{\eta}\sim10^{11}\,{\rm GeV}$. Interestingly, the constraint from proton decay (see Sec.~\ref{sec:observable}) ensures that this condition is satisfied {\em automatically}. The primordial lepton asymmetry, when reprocessed by sphaleron interactions, leads to a similar baryon asymmetry \cite{CDO} \begin{equation} {n_B\over n_\gamma}\sim{n_L\over n_\gamma}\sim \left({m_{\eta}\over M_{Pl}}\right)^{1/2}\epsilon\sim 10^{-4}\ \epsilon\ , \label{eq:nB} \end{equation} where the asymmetry parameter ($\epsilon$) due to the decay of the $i$th-generation neutrino and sneutrino is given by \cite{CDO} \begin{equation} \epsilon_i = {1\over 2\pi (\lambda^\dagger_3\lambda_3)_{ii}} \sum_j \left({\rm Im}\,[(\lambda^\dagger_3\lambda_3)_{ij}]^2\right)\, g(M^2_{\nu^c_j}/M^2_{\nu^c_i})\ , \label{eq:eps} \end{equation} with \begin{equation} g(x)=4\sqrt{x}\,\ln{1+x\over x}\ . \label{eq:g} \end{equation} To proceed we need to manipulate the entries in $\lambda_3$, which has remained as yet unspecified. We define the unitary rotation matrix $U$, such that $\hat\lambda_3=U\lambda_3 U^\dagger$, where $\hat\lambda_3$ is the diagonal matrix of eigenvalues of $\lambda_3$. Experience with the quark mixing matrix leads us to assume that $U$ differs little from the identity matrix: $U={\bf1}+R$, with \cite{ELNO} \begin{equation} R=\left(\begin{array}{ccc} 0&\theta_{12}&0\\ -\theta^*_{12}&0&\theta_{23}\\ 0&-\theta^*_{23}&0\end{array}\right)\ . \label{eq:R} \end{equation} With this ansatz we obtain to lowest non-vanishing order \begin{eqnarray} (\lambda^\dagger_3\lambda_3)_{ii}&=&|\hat\lambda^i_3|^2 +\sum_j|\hat\lambda^j_3|^2\, |\theta_{ij}|^2 \ , \label{eq:ii}\\ (\lambda^\dagger_3\lambda_3)_{ij}&=&|\theta_{ij}|e^{i\phi_{ij}} \left[|\hat\lambda^i_3|^2-|\hat\lambda^j_3|^2\right]\quad (i\not=j)\ , \label{eq:ij} \end{eqnarray} where $\phi_{ij}={\rm Arg}\,[\theta_{ij}]$. Thus, Eq.~(\ref{eq:eps}) becomes \begin{equation} \epsilon_i={1\over2\pi}\, {\sum_{j\not=i}|\theta_{ij}|^2\sin2\phi_{ij}\, [|\hat\lambda^i_3|^2-|\hat\lambda^j_3|^2]^2\ g(M^2_{\nu^c_j}/M^2_{\nu^c_i})\over |\hat\lambda^i_3|^2+\sum_j|\hat\lambda^j_3|^2\,|\theta_{ij}|^2}\ . \label{eq:eps2} \end{equation} Because of the several unknown parameters in the above expressions, and the inherent uncertainties in this type of calculations, we will be content with presenting a plausible scenario leading to interesting lepton asymmetries and see-saw neutrino masses. For simplicity let us assume that the $\lambda_6$ matrix is proportional to the unit matrix, {\em i.e.}, \begin{equation} M_{\nu^c_1}=M_{\nu^c_2}=M_{\nu^c_3}=M_{\nu^c} =\lambda_6\,{\vev{T\bar T}\over M}\ . \label{eq:Mnu^c} \end{equation} The light neutrino mass matrix then becomes $M_\nu=\lambda_3\lambda^T_3 v^2_2/M_{\nu^c}$. If we neglect the CP violating phases (a not necessarily justified approximation), the matrix $U$ which diagonalizes $\lambda_3\lambda^\dagger_3$, also diagonalizes $\lambda_3\lambda^T_3$ and the physical neutrino masses become (up renormalization group scaling corrections \cite{chorus}) \begin{equation} m_{\nu_i}\approx{(\hat\lambda^i_3 v_2)^2\over M_{\nu^c}}\ . \label{eq:Mnu} \end{equation} Furthermore, in our ansatz the (small) neutrino mixing angles are given by $\theta_{e\mu}=\theta_{12}$, $\theta_{e\tau}=0$, and $\theta_{\mu\tau}=\theta_{23}$. As we will see shortly, these mixing angles are unrestricted from lepton asymmetry considerations, and thus could accomodate the MSW solution to the solar neutrino problem ($\nu_e\leftrightarrow\nu_\mu$) and lead to interesting $\nu_\mu\leftrightarrow\nu_\tau$ oscillations at the CHORUS and NOMAD, and P803 experiments at CERN and Fermilab respectively. {}From Eq.~(\ref{eq:Mnu}) we see that $m_{\nu_\tau}\approx (\hat\lambda_3^3 v_2)^2/M_{\nu^c} =[\hat\lambda_3^3\, \sin\beta(174\,{\rm GeV})]^2/M_{\nu^c}$. With $M_{\nu^c}\sim10^{10}\,{\rm GeV}$ and $\hat\lambda^3_3\approx0.1$, we get $m_{\nu_\tau}\sim15\,(30)\,{\rm eV}$ for $\tan\beta\sim1\,(\tan\beta\gg1)$. This range of tau neutrino masses provide an adequate and desirable hot dark matter component of the Universe. Thus, in what follows we take $\hat\lambda^3_3=0.1$. It is also natural to assume that the remaining eigenvalues of the $\lambda_3$ matrix are hierarchically smaller, {\em i.e.}, $\hat\lambda^1_3\ll\hat\lambda^2_3\ll\hat\lambda^3_3$. For instance, $\hat\lambda^2_3\sim{1\over100} \hat\lambda^3_3$ yields $m_{\nu_\mu}\sim 10^{-3}\,{\rm eV}$, consistent with solutions to the solar neutrino problem via the MSW mechanism. (These hierarchies are comparable to those in the up-quark Yukawa matrix.) Going back to the calculation of the lepton asymmetries, with our hierarchical assumption for $\hat\lambda_3$, from Eq.~(\ref{eq:eps2}) we obtain \begin{eqnarray} \epsilon_1&\approx&{2\ln2\over\pi}\,\left(\hat\lambda^2_3\right)^2 \sin2\phi_{12}\sim 10^{-6}\,\phi_{12}\ , \label{eq:e1}\\ \epsilon_2&\approx&{2\ln2\over\pi}\,\left(\hat\lambda^3_3\right)^2 \sin2\phi_{23}\sim 10^{-2}\,\phi_{23}\ , \label{eq:e2}\\ \epsilon_3&\approx&{2\ln2\over\pi}\,\left(\hat\lambda^3_3\right)^2 |\theta_{23}|^2\sin2\phi_{23}\sim 10^{-2}\,|\theta_{23}|^2\phi_{23}\ . \label{eq:e3} \end{eqnarray} With the expression for the estimated baryon asymmetry in Eq.~(\ref{eq:nB}), we would get the desired result of ${\rm few}\times10^{-10}$ for $\phi_{12}\sim1$ and $\phi_{23}\ll1$. The natural choice would be maximal CP violation in the $\theta_{12}$ entry in the rotation matrix $R$ (see Eq.~(\ref{eq:R})) and no CP violation elsewhere in the matrix (unless new entropy diluting sources are introduced to reduce $\epsilon_1+\epsilon_2+\epsilon_3$). These results would be affected somewhat if one allows a non-trivial structure to the matrix $\lambda_6$ ({\em i.e.}, relaxing the assumption in Eq.~(\ref{eq:Mnu^c})). We now remark that this model is also viable in the traditional out-of-equilibrium scenario, where $\epsilon_1$ is the only surviving asymmetry. The out-of-equilibrium condition at $T=M_{\nu^c_1}\sim10^{10}\,{\rm GeV}$, \begin{equation} \Gamma_{\nu^c_1}={(\lambda^\dagger_3\lambda_3)_{11}\over16\pi}\, M_{\nu^c_1}<1.66 g^{1/2}_*\,{T^2\over M_{Pl}}=H\ , \label{eq:out} \end{equation} is satisfied for (using Eq.~(\ref{eq:ii})) \begin{equation} (\lambda^\dagger_3\lambda_3)_{11}= |\hat\lambda^1_3|^2+|\hat\lambda^2_3|^2\,|\theta_{12}|^2\lsim10^{-6}\ , \label{eq:cond} \end{equation} which is consistent with our hierarchical assumption. However, in this case the calculation of the leptonic asymmetry has a larger ($\sim10^{-2}$ \cite{CDO}) coefficient than in Eq.~(\ref{eq:nB}), requiring a non-maximal CP violating phase $\phi_{12}\sim10^{-2}$. Finally, let us comment on whether or not the sphaleron interactions may wash away the leptonic asymmetry produced above. This could in principle happen if the non-renormalizable operators obtained when integrating out the right-handed neutrino fields, {\em i.e.}, $(\lambda_3/M_{\nu^c}) LLHH$, where $L$ is the lepton doublet in $\bar f$ and $H$ the Higgs doublet in $h$, are in equilibrium with the sphaleron interactions \cite{FY}. It has been shown \cite{CKO} that to prevent the erasure of the asymmetry, one must demand $M_{\nu^c}\mathrel{\mathpalette\@versim>}(\lambda_3)^2\, 3\times10^9\,{\rm GeV}$, which is always satisfied for our choices of $\lambda_3$ and $M_{\nu^c}$. \section{Comparison with $SU(5)\times U(1)$} \label{sec:comparison} The revamped MDM presented in the previous sections has several appealing phenomenological features, constituting an interesting example of traditional grand unified model building. Nonetheless, it is apparent that the model is rather non-minimal or uneconomical. For instance, a \r{75} needs to be used for GUT symmetry breaking, greatly increasing the size of the GUT particle spectrum. Moreover, the \r{50},\rb{50} to effect the doublet-triplet splitting problem make the unified gauge coupling so large above the GUT scale that they need to be taken at the gravitational scale. The doublet-triplet splitting is tamed, but proton decay can still be too fast because of the ``useless" pieces of the \r{50},\rb{50} representations which need to be made heavy, resulting in the otherwise-not-needed doubling of these representations and of the Higgs pentaplets. Regarding the right-handed neutrinos, their (ad-hoc) introduction has various desirable consequences. However, the Yukawa matrix coupling them to the lepton doublets is arbitrary, with no particular motivation for its desired hierarchical structure. It is interesting to note that the above critique of the revamped MDM can be circumvented altogether if one extends the gauge group from $SU(5)$ to $SU(5)\times U(1)$ \cite{Barr,revitalized,Moscow}. Gauge symmetry breaking down to the Standard Model gauge group occurs via vacuum expectation values of the $H$ (\r{10}) and $\bar H$ (\rb{10}) Higgs representations. This is possible because of the ``flipping" $u\leftrightarrow d$, $u^c\leftrightarrow d^c$, $e\leftrightarrow\nu$, $e^c\leftrightarrow\nu^c$ involved in the assignment of the Standard Model particles to the $\bar f=\{u^c,L\}$ (\rb{5}) and $F=\{Q,d^c,\nu^c\}$ (\r{10}) representations. Thus, $H$ and $\bar H$ contain one pair of neutral fields $\nu^c_H,\nu^c_{\bar H}$, which get vevs along the flat direction $\vev{\nu^c_H}=\vev{\nu^c_{\bar H}}$. There is no need for large GUT representations for symmetry breaking. As is well known (and we review below), this property takes on a much larger magnitude when one attempts to derive these models in string model building. The missing-partner mechanism, which above involved the couplings $\bar\theta\Sigma h$ [(\rb{50})(\r{75})(\r{5})] and $\theta\Sigma\bar h$ [(\r{50})(\r{75})(\rb{5})], is now effected by the couplings $HHh$ [(\r{10})(\r{10})(\r{5})] and $\bar H\bar H\bar h$ [(\rb{10})(\rb{10})(\rb{5})]. First note that no additional representations are needed besides the GUT-breaking Higgs ones. Moreover, the resulting Higgs triplet matrix \begin{equation} \bordermatrix{ &\bar h_3&d^c_H\cr h_3&0&\lambda_4\,\vev{\nu^c_H}\cr d^c_{\bar H}&\lambda_5\,\vev{\nu^c_{\bar H}}&0}\ , \label{eq:f2/3} \end{equation} does not need a large non-zero (22) entry ({\em c.f.} Eq.~(\ref{eq:2/3})) because the ``useless" components of the $H$ and $\bar H$ representations are eaten by the GUT gauge bosons to become massive or become GUT Higgsinos. This natural zero mass term for $d^c_H d^c_{\bar H}$ implies that the dimension-five proton decay operators are negligible. We end up with a very economical GUT Higgs spectrum and no threat of dimension-five operators. Regarding neutrino masses, the right-handed neutrinos which had to be introduced by hand in the revamped MDM, are now contained in the $F$ (\r{10}) representations. Indeed, the coupling $\lambda_3\bar f\nu^c h$ in Eq.~(\ref{eq:W'''}) is here written as $\lambda_3\bar f e^c h$, with the (unavoidable) right-handed electrons now introduced ``by hand". In $SU(5)\times U(1)$ this coupling provides the charged lepton masses. On the other hand, the coupling $\lambda_1 F\bar f\,\bar h$, which in Eq.~(\ref{eq:W}) provided the down-quark masses, here provides the up-quark masses and Dirac neutrino masses. (Also, the coupling $\lambda_2 FFh$, which in Eq.~(\ref{eq:W}) provided the up-quark masses, here provides the down-quark masses.) Thus, the right-handed neutrinos are unavoidable in $SU(5)\times U(1)$, and their Yukawa couplings to the lepton doublets are equal to those of the up-quark Yukawa matrix, providing (as discussed in Sec.~\ref{sec:CBA}) an automatic and desirable hierarchy in the see-saw neutrino masses. An important distinction between the see-saw mechanism in the revamped MDM and $SU(5)\times U(1)$ is the manner in which the right-handed neutrinos get a mass. In the revamped MDM this is through the superpotential term $\lambda_6\nu^c\nu^c\vev{T\bar T}/M$ in Eq.~(\ref{eq:W'''}), whereas in $SU(5)\times U(1)$ there are two possible sources: (i) through cubic couplings $\lambda_6 F\bar H\phi\ni\lambda_6\vev{\nu^c_{\bar H}}\nu^c\phi$, where $\phi$ (with $\vev{\phi}=0$) are $SU(5)$ singlets \cite{revitalized}; and (ii) through non-renormalizable couplings $\lambda_9 FF\bar H\bar H/M\ni \lambda_9 (\vev{\nu^c_{\bar H}}^2/M)\nu^c\nu^c$ \cite{improved}. The second form resembles that in the revamped MDM, although the mass scale is likely higher ({\em i.e.}, $\vev{\nu^c_{\bar H}}^2/M\sim10^{14}\,{\rm GeV}$). These two models also differ somewhat in the calculation of the cosmic baryon asymmetry, besides the possible difference in the right-handed neutrino mass spectrum. Indeed, because the $SU(5)\times U(1)$ gauge symmetry is broken along a flat direction, there is a dilution factor ($\Delta$) in the computation of the lepton asymmetry due to the entropy released by the late-decaying ``flaton" field \cite{dilution,ENO}. However, these two effects ($\nu^c$ spectrum and $\Delta$) tend to compensate each other and an acceptable baryon asymmetry is typically obtained \cite{ENO,ELNO}. There is another cosmological aspect of these models that sets them apart, namely the breaking of the GUT symmetry down to the Standard Model one. In the MDM, $SU(5)$ symmetry breaking via an arbitrary vev of the \r{75} leads to several possible degenerate vacua \cite{HMPR}, at least in the context of global supersymmetry. When supergravity effects are taken into account, if the desired vacuum has zero cosmological constant, all the others will be lower in energy, although essentially unreachable \cite{Weinberg}. Thus, if the vev of the \r{75} can be arranged to be in the desired direction, the Universe will remain in the desired broken phase. In contrast, in $SU(5)\times U(1)$ the breaking down to the Standard Model via the vevs of the \r{10},\rb{10} along the F- and D-flat direction $\vev{\nu^c_H}=\vev{\nu^c_{\bar H}}$ is {\em unique} \cite{revitalized}. Regarding the issue of unification, the revamped MDM requires non-minimal representations to make this possible. In $SU(5)\times U(1)$ traditional grand unification does not occur (although the non-abelian Standard Model gauge groups do unify) and unification is not a test of the model. However, if string unification is desired (at the scale $M_{\rm str}\approx4\times10^{17}\,{\rm GeV}$), then non-minimal representations need to be added to the $SU(5)\times U(1)$ model \cite{gap}. We have seen that the pair of \r{50},\rb{50} representations in the revamped MDM need to be put at the gravitational scale. It is then natural to ask whether this model can be obtained from the only known consistent theory of quantum gravity, namely string theory. Because of some technical difficulties which we review below, no attempts have been made to derive the MDM from strings. It is of course well known that $SU(5)\times U(1)$ can be easily derived from strings \cite{revamp,search}. The prime constraint in string model-building is that of the massless representations which are allowed when the corresponding gauge group $G$ is represented by a ``level-$k$" Kac-Moody algebra on the world-sheet \cite{ELN,GO}. The allowed representations must be unitary, \begin{equation} \sum_{i=1}^{\rm rank\, G} n_i m_i\le k\ , \label{eq:unitary} \end{equation} where $n_i$ are the Dynkin labels of the highest weight representation in question, and $m_i$ are fixed positive integers for a given $G$. In the case of $SU(n)$: $m_i=1\,,\,\forall i$. In our $SU(5)$ example then $\sum_{i=1}^4 n_i\le k$. Looking up the $n_i$ values, we see that for $k=1$, only \r{1},\r{5},\rb{5},\r{10},\rb{10} are unitary. For $k=2$ we find in addition: \r{15},\r{24},\r{40},\rb{40},\r{45},\rb{45},\r{50},\rb{50},\r{75}. Only level-one constructions appear to be needed to derive $SU(5)\times U(1)$, whereas at least level-two constructions are required in the MDM. However, this is not the end of the story, since one can also ask whether the allowed representations could possible be massless. This requires calculating the so-called conformal dimension $h_r$ of the representation $r$, \begin{equation} h_r={C_r\over 2k+C_A}\ , \label{eq:h} \end{equation} where $C_r$ is the Casimir of $r$, and $C_A$ that of the adjoint representation. If $h_r>1$, the representation is necessarily massive. For $h_r\le1$ the representation could be massless, although this is not guaranteed since other degrees of freedom may add their own contribution to the conformal dimension making it exceed unity. It is not hard to see that in $SU(5)$ all unitary representations at level one are also massless \cite{ELN}, and thus $SU(5)\times U(1)$ models can be readily constructed at level one. The unitary representations of interest for MDM model-building, which are allowed at level two, have conformal dimensions \begin{equation} h_{50,\overline{50}}={42\over5(k+5)}\ ,\qquad h_{75}={8\over k+5}\ , \label{eq:h5075} \end{equation} and are not massless at level two. In fact, $k=4$ is required to make all these representations massless. Such high-level Kac-Moody constructions have never been attempted. One intriguing possibility would be to construct level-two $SU(5)$ string models (for recent attempts see Ref.~\cite{Aldazabal}), which should allow the required large MDM representations, although with masses at the Planck scale. Note that this is not necessarily a problem since we already require the \r{50},\rb{50} to be at that mass scale. If the \r{75} is also raised to that scale, the breaking of $SU(5)$ would occur at the string scale, and this may be difficult to reconcile with gauge coupling unification. It is also worth remarking that in a string model all gauge couplings are related at the string scale, and with $SU(5)$ constructed at level two, the relation would be $\sqrt{2}\,g_5=g_h$ \cite{Ginsparg}, where $g_h$ is the gauge coupling of the hidden gauge group. Finally, the mass terms in Eqs.~(\ref{eq:W},\ref{eq:W'}), which would not be allowed if the large MDM representations belonged to the massless spectrum, are expected to arise when they belong to the string massive spectrum. Of course, bridging the gap between the massless and massive spectrum may create problems in obtaining the low-energy effective field theory, but this question cannot be answered until an actual model is constructed along these lines. \section{Conclusions} \label{sec:conclusions} During the last few years, a great deal of attention has been paid to supersymmetric grand unified theories in light of the precise LEP measurements of the Standard Model gauge couplings. These analyses depend crucially on the details of the low-energy supersymmetric spectrum and the heavy GUT spectrum. Most of the effort to date has been focused on the minimal $SU(5)$ supergravity model, which appears to be running into difficulties regarding unification and proton decay. In addition, there is the nagging doublet-triplet splitting problem that receives no satisfactory explanation. Motivated by these developments, we have reconsidered one of the alternatives to minimal $SU(5)$, where the doublet-triplet splitting is dealt with in a reasonable way via the missing-partner mechanism, and gauge coupling unification is not in jeopardy. We have revamped this model to tame the dimension-five proton decay operators, and to allow see-saw neutrino masses. In order to generate the needed intermediate scale for the right-handed neutrino masses, we have endowed the model with a ``modern" hidden sector which can generate dynamically the desired intermediate scale, the scale of supersymmetry breaking, and the Higgs mixing parameter $\mu$. The revamped MDM also provides for the cosmic baryon asymmetry through the Fukugita-Yanagida mechanism via lepton-number-violating decays of the right-handed neutrinos. We have also contrasted the main features of the revamped MDM against the ``flipped" $SU(5)\times U(1)$ model, and basically shown that the former can be considered as a ``poor man's" version of the latter. In the realm of string model-building, $SU(5)\times U(1)$ fares rather well, whereas the revamped MDM is very unlikely to be realized, except perhaps if one allows $SU(5)$ symmetry breaking to occur at the string scale. \section*{Acknowledgments} We would like to thank John Ellis and Bruce Campbell for useful discussions. This work has been supported in part by DOE grant DE-FG05-91-ER-40633.
\section{Introduction} \label{sec-intro} Remarkable, exact results on the vacuum structure of four dimensional $N=2$ supersymmetric quantum field theories with an $SU(2)$ gauge symmetry and matter fields in the fundamental representation have recently been obtained \cite{SW}, with generalizations to $SU(N_c)$ gauge groups \cite{Ncolor}. In these models, an $N=1$ supersymmetry preserving mass perturbation for the adjoint superfield leads to monopole condensation, confinement and chiral symmetry breaking. There has been much interest in how such exact supersymmetric results generalize to nonsupersymmetric gauge theories \cite{othersusybreaking}. In this paper we show how $N=2$, $N=1$ and softly broken supersymmetric models can all be embedded in a single enlarged $N=1$ model in which coupling constants are treated as ``spurion'' fields. The effective superpotential of the enlarged model can be exactly determined from its symmetries and the $N=2$ limit. In the $N=2$ limit the $D$-terms are also exactly determined, but when supersymmetry is broken the potential and superpotential each receive contributions from additional $D$-terms of the enlarged model, which vanish in the $N=2$ limit. Although constrained by the symmetries of the model, these cannot be fully determined. If the $N=2$ symmetry is broken solely by soft breaking masses, these unknown functions typically are necessary to determine details of the vacuum structure. The exact superpotential does however include all the lowest-dimension gauge interaction terms, and continues to have singularity structure\footnote{ Strictly speaking there is no longer a moduli space after breaking $N=2$ supersymmetry. However, we can still discuss monodromies in the configuration space of the theory, by means of an external source used to traverse a closed loop.} consistent with points at which monopoles become massless. We can thus confirm that when the soft breaking masses are perturbations to an $N=1$ preserving mass, the vacuum remains essentially that of the $N=1$ model. An important step in obtaining the exact results of \cite{SW} is to treat the couplings in the $N=2$ models as spurion chiral superfields in an enlarged $N=1$ model, which fail to propagate if their kinetic term coefficients are taken to infinity \cite{natiral}. Their scalar component vacuum expectation values (vevs) can thus be frozen at any chosen values. The $F$-terms of the Wilsonian effective Lagrangian of a supersymmetric model \cite{natiral,SV,CERN} must be holomorphic in the fields as well as invariant under the gauge and global symmetries of the model (assuming a supersymmetric regularization scheme). The superpotential of the enlarged model is therefore holomorphic in the couplings. A recent paper \cite{yale1} showed that soft supersymmetry breaking interactions and mass terms can be introduced in supersymmetric models without altering the holomorphic constraint on the $F$-terms. The spurion fields can be coupled to a sector that generates supersymmetry breaking expectation values for the spurion field $f$-components\footnote{ Our superfield notation \cite{WB} is $\Phi = a_\Phi + \sqrt{2}\, \theta \psi_\Phi + \theta \theta f_\Phi$, with $\Phi^\dagger\Phi\big|_D \equiv \int d^2\theta\ d^2{\bar\theta}\ \Phi^\dagger\Phi = |f_\Phi|^2,\ \Phi\big|_F \equiv \int d^2\theta\ \Phi = f_\Phi,\ \Phi\big|_A \equiv a_\Phi$, etc.}. Freezing out the spurions generates soft supersymmetry breaking masses and interactions in the embedded model. For large vevs of the adjoint scalar field $a_\Phi$ in the models we study, the $SU(N_c)$ gauge symmetry is broken to a subgroup of weakly interacting $U(1)$ gauge symmetries. It has been argued \cite{SW} that the $N=2$ theory remains in the Coulomb phase even for small scalar vev where the model is strongly coupled. One might have expected the pure $SU(N_c)$ dynamics to operate unimpeded when the scalar vev was much smaller than the strong interaction scale $\Lambda$, and the low energy theory to exist in a confining phase with all the gauge bosons confined within glueballs. However, the solutions' self-consistency \cite{SW} suggests that for the special choice of couplings that gives the $N=2$ model, the quantum effects that cause condensation and confinement exhibit special cancellations, leading to the surprising low energy dynamics of a strongly coupled Coulomb phase with massless fermions, monopoles and dyons. The solutions' consistency with our expectations for how these theories ought to behave helps confirm that higher dimension terms do not qualitatively change the solutions. Following the analysis \cite{SW}, which does not determine operators with more than two derivatives (${\cal O}(p^2)$) or four fermions, we will assume that the gauge particle dynamics are essentially controlled by the effective operators of lowest dimension, at sufficiently low energy scales. The model's symmetries permit higher dimension operators containing for example powers of $|WW/\Lambda^3|^2$, and (by $N=2$ supersymmetry) also containing superderivatives and powers of $\Phi/\Lambda$. Since the scalar component of $\Phi$ is pinned at $\langle a_\Phi \rangle \sim \Lambda$ when the supersymmetry is broken to $N=1$, these terms are not at first sight negligible. However, since they must be related by $N=2$ supersymmetry to the corresponding higher derivative $W$ terms \cite{mans}, they should also be suppressed in the infrared regime. Thus despite these caveats, a tractable electric--magnetic duality is useful in determining the theory's vacuum structure. We expect that introducing $N=2$ breaking masses destroys cancellations in the dynamics below the scale of the breaking mass, and confines the $U(1)$ gauge interaction below that scale (via the Higgs mechanism in the dual theory). Provided that breaking masses are sufficiently small relative to the scale $\Lambda$ of the strong interactions, the theory will continue to be described by a $U(1)$ gauge theory between the $SU(2)$ breaking scale parameterized by $\langle a_\Phi \rangle$, pinned at $\sim \Lambda$, and the breaking mass scale. Softly broken $N=1$ and $N=0$ models, when close in parameter space to the $N=2$ model, are therefore anticipated to occupy the same phase. The paper is organized as follows. Section~\ref{sec-su2} explicitly describes the class of models we study. We allow the models' coupling constants to be chiral superfield spurions, whose values are eventually frozen in all of spacetime \cite{natiral,yale1}. We then review the analysis that leads to the Seiberg--Witten ansatz for the vacuum structure for $SU(2)$ gauge symmetry, and monopole condensation in $N=1$ models close in parameter space to the $N=2$ model. Allowing soft supersymmetry breaking terms in our Lagrangian via nonvanishing spurion field $f$-components, we show that to lowest order in the soft breakings, gauge kinetic terms are unchanged by the introduction of soft masses. However, the potential minimum in these models typically depends upon insufficiently constrained contributions from $D$-terms. Section~\ref{sec-matter} discusses obtaining similar results in $SU(N_c)$ gauge theories or with matter fields in the fundamental representation, and reaching QCD by continuous interpolation from models of this type. \newpage \section{Supersymmetric $SU(2)$} \label{sec-su2} \subsection{An enlarged model} \label{sec-2.1} Consider an $N=1$ supersymmetric $SU(2)$ gauge theory, containing one matter chiral superfield in the adjoint representation. Promoting the coupling constants into chiral superfields (including a $D$-term normalization, $K$, which we shall use to generate a squark mass) \cite{yale1} yields the Lagrangian \begin{equation} \label{smallLagrangian} \begin{array}{c} {\cal L} = {1\over 4\pi}{\rm Im}\left( {1\over 2}\tau_0 W_\alpha W^\alpha \Big|_F \ +\ (\tau_0 + K^\dagger K) \Phi^\dagger e^V \Phi\Big|_D \right) \ +\ m \Phi^2\Big|_F \ +\ {\rm h.c.}\\ \\ \ +\ \Lambda_m^2 \left(m^\dagger m\Big|_D + \beta_m m\Big|_F + {\rm h.c.}\right) \ +\ \Lambda_\tau^2 \left( \tau_0^\dagger \tau_0 \Big|_D + \beta_\tau \tau_0 \Big|_F + {\rm h.c.}\right)\\ \\ \ +\ \Lambda_K^2 \left( K^\dagger K\Big|_D + \beta_K K\Big|_F + {\rm h.c.}\right) \end{array} \end{equation} (note that $\Lambda_m$ and $K$ are dimensionless). The spurion fields $i = \{m, \tau_0, K\}$ do not propagate in the limit $\Lambda_i \rightarrow \infty$; we can fix their scalar components $a_i$ at any expectation values we choose, and replace $f_i = -\beta_i^\dagger + {\cal O}(1/\Lambda_i^2)$. [An equivalent approach normalizes the spurion $D$-terms conventionally, placing the $\Lambda_i$ suppression on the spurion-to-physical-field coupling operators; for example writing $m^\dagger m\Big|_D + (m/\Lambda_m)\Phi^2\Big|_F$. The spurion vevs are then taken proportional to $\Lambda_i$, so their classical and quantum evolutions are still relatively suppressed.] In any case, if we choose $\langle a_{\tau_0} \rangle\neq 0$ and $\left(\langle a_m\rangle\ ,\ \langle a_{K} \rangle\ ,\ \langle f_i \rangle\ \right) \ =\ 0$, the model reduces to $N=2$ supersymmetric $SU(2)$, studied in \cite{SW}, whose bare Lagrangian is \begin{equation} \label{neqone} {\cal L}_{N=2}\ =\ {1\over 4\pi}{\rm Im} \langle a_{\tau_0}\rangle \left( {1\over 2} W_\alpha W^\alpha \Big|_F \ +\ \Phi^{\dagger} e^V \Phi\Big|_D \right)\ . \end{equation} In this notation, we will review the derivation \cite{SW} of the $N=2$ model effective potential and its generalization to the $N=1$ case of $\langle a_m \rangle \neq 0$. Nonzero values of $\langle f_{\tau_0} \rangle$ and $\langle f_K\rangle$ generate supersymmetry breaking mass terms for gauginos and squarks respectively. In the limit $\langle f_{\tau_0,K} \rangle \rightarrow\infty$ the gauginos and squarks decouple from the low energy theory, leaving only $SU(2)$ gauge bosons and adjoint fermions. The model has an anomaly-free $U(1)_R$ global symmetry, whose charge assignments \begin{equation} \label{U1Rcharges} \matrix{ \theta && +1 \cr W_\alpha && +1 \cr \tau_0 && 0 \cr \Phi && 0 \cr m && +2 \cr K && {\rm arbitrary} } \end{equation} significantly constrain the terms which can appear in the low energy effective Lagrangian. Allowable $D$-term operators have net $R$-charge of zero, whereas $F$-term operators must have net $R$-charge of $+2$. The anomaly breaks a second $SU(2)_R$ symmetry of the embedded $N=2$ model to a discrete symmetry; in the $SU(2)$ gauge case this gives the ${\bf Z_2}$ symmetry. A further constraint applies to the spurion sources $\beta_i$, which may be treated as decoupled chiral superfields in their own right. Coupled only linearly to the $f$-components of $m,\tau$ and $K$, they do not contribute perturbatively to 1PI diagrams. Moreover, since scalar components $a_{\beta_i}$ do not couple at all to other fields in the bare potential, they will only appear suppressed by powers of $\Lambda_i$ in induced terms in the effective Lagrangian. The only dependence in the effective theory on $\beta_i$ vevs thus occurs indirectly, through the $f$-component vevs $f_i$ of coupling-field spurions, induced by $a_{\beta_i}$. \subsection{The low energy effective theory in the $N=2$ limit} \label{sec-2.2} To fix our notation we briefly review the analysis \cite{SW} relevant to the $N=2$ model in (\ref{neqone}). The form of the ${\cal O}(p^2)$ effective Lagrangian's superpotential can be deduced from its holomorphic properties as described below. In the $N=2$ limit the effective Lagrangian's $D$-terms are then determined by $N=2$ supersymmetry, giving the effective Lagrangian as \begin{equation} \label{N=2lowE} {1\over 4\pi} {\rm Im}\left\{ {\partial{\cal F}(A) \over \partial A} \bar A\Bigg|_D \ +\ {1\over 2} {\partial^2{\cal F}(A) \over \partial A^2} W_{\alpha} W^{\alpha}\Bigg|_F \right\} \ , \end{equation} where $A$ is the $N=1$ chiral multiplet and ${\cal F}$ the ``prepotential'' \cite{SW,prepotential}. The exact solution for the effective Lagrangian of the $N=2$ model determines the effective Lagrangian of the enlarged model, and hence of the models with soft supersymmetry breaking masses, up to terms compatible with the symmetries of the enlarged model that vanish in the $N=2$ limit. We shall discuss these extra terms in the following sections. In the $N=2$ model the classical potential for $a_\Phi$, $V_{cl} = {\rm Tr}([a_\Phi, a_\Phi^\dagger]^2)$, is minimized (and vanishes) when $a_\Phi$ has any diagonal complex vev. This breaks $SU(2)$ down to $U(1)$ and yields a classical moduli space, parameterized by the gauge invariant superfield $U \equiv {\rm Tr}[\Phi^2]$. We assume that at low energies the quantum theory also remains in the Coulomb phase, so that the particle content is restricted to the $U(1)$ gauge superfield $W_\alpha$ and the massless matter fields. In the Coulomb phase at arbitrarily low scales, where we need consider only the lowest dimension terms in the Lagrangian, the effective action is simply quadratic in the gauge field, and a duality transformation is conveniently derived \cite{SW} by considering a functional integral over the gauge superfield $W$. Imposing the condition Im$({\cal D} W) = 0$ via a Lagrange multiplier $V$, incorporated by a new dual gauge field $W_D = i{\cal D}V$, we can complete the square and perform the Gaussian integral, to obtain an equivalent theory for the dual field with $\tau_D = -1/\tau$: \begin{equation} {\cal Z} \sim \int{\cal D}[W_D]\ \exp{\left({i\over 8\pi} {\rm Im} \int \tau_D W_D^\alpha W_{D \alpha}\Big|_F \right)}\ . \end{equation} In order to maintain the explicit $N=2$ supersymmetry of (\ref{N=2lowE}) in the dual theory we must rewrite the matter $D$-terms \begin{equation} \label{dualprepotential} {\partial {\cal F}(A) \over \partial A} \bar{A} \ =\ {\partial {\cal F_D}(A_D) \over \partial A_D} \bar{A}_D \ , \end{equation} with $\tau \ =\ \partial^2 {\cal F}(A)/ \partial A^2$ and $\tau_D \ =\ \partial^2 {\cal F}_D(A_D)/ \partial A_D^2$. Thus \begin{equation} \label{taudfda} A_D \ =\ {\partial {\cal F(A)} \over \partial A}\ , \qquad \tau \ =\ {\partial A_D \over \partial A}\ . \end{equation} This manipulation is independent of whether $\tau$ is a coupling constant or, as in our enlarged model, a chiral superfield itself. Duality therefore continues to hold at low energies even after supersymmetry is broken by for example $\langle f_\tau \rangle \neq 0$, induced by $f_{\tau_0} \neq 0$; in the dual theory, the trivially related $\langle f_{\tau_D} \rangle$ drives supersymmetry breaking. The theory is also invariant under shifts by integer $n$ in the real part of $\tau$, since these correspond to unobservable shifts in the $\theta$ angle. Writing $\tau$ as the ratio of the two components of a vector allows us to recognize the duality and $\theta$ angle transformations \begin{equation} \label{tautransform} \left( \begin{array}{cc} a & b\\ c & d \end{array} \right) \left( \begin{array}{c} \tau \\ 1 \end{array} \right) \end{equation} as generating the group $SL(2,Z)$. Now consider the effective theory for scales much less than $U$. Its $F$-terms must be holomorphic and invariant under the global $U(1)_R$ symmetry of (\ref{U1Rcharges}), and therefore of the form \begin{equation} {\cal L}_{\rm gauge} \ =\ {1\over 8\pi}{\rm Im} \left[ \tau(\tau_0,U)\, W_\alpha W^\alpha \Big|_F \right] \ . \end{equation} If $\tau$ were everywhere analytic in $U$, Im$(\tau)$ would be a harmonic function, which would be unbounded below, resulting in an imaginary gauge coupling. To avoid this $\tau$ must have singularities at finite values of $U$, $U_i$, presumably due to composite states driven massless by strong interactions. Thus the dual theory is weakly coupled near the singularities. The composite (monopole or dyon) fields are light near the singular points and must be included in the effective theory. We therefore include ``hypermultiplet'' terms compatible with the symmetries in the effective Lagrangian: \begin{equation} \label{Lmonopole} {\cal L}_{\rm dyon} \ =\ M^\dagger e^V M \Big|_D \ +\ {\tilde M}^\dagger e^{-V} {\tilde M} \Big|_D \ +\ \sqrt{2} A_D M {\tilde M} \Big|_F \ \ +\ {\rm h.c.}\ , \end{equation} where $M$ and ${\tilde M}$ are the two $N=1$ chiral multiplets of an $N=2$ hypermultiplet, and $A_D \sim (U-U_i)$ determines the bound states' mass. The simplest possibility for $\tau$, consistent with its weak-coupling limit at large $U$ and with ${\bf Z_2}$ symmetry $U \leftrightarrow -U$, is a pair of logarithmic singular points at each of which a single composite state becomes massless. An $SL(2,Z)$ invariant function with two such singularities is determined up to scalings by its behavior there and at infinity, and can be interpreted as the modular parameter of a torus (with $\tau$ the ratio of its two periods). The torus described by $\tau$ corresponds to a cubic elliptic curve with roots $e_1, e_2, e_3$, \begin{equation} \label{ellcurve} y^2 \ =\ 4 (x - e_1) (x - e_2) (x - e_3) \ . \end{equation} Up to $SL(2,Z)$ transformations, $\tau$ can be written \begin{equation} \label{tau} \tau \ =\ {\partial A_D \over \partial A} \ =\ {\int_{e_1}^{e_2} {dx / y} \over \int_{e_2}^{e_3} {dx / y}}\ . \end{equation} Seiberg and Witten showed that if the dyons are respectively a $(0,1)$ magnetic monopole and a $(1,-1)$ dyon, then one obtains a function $\tau$ consistent with the one loop beta function at $U \rightarrow \infty$: $\tau(U\rightarrow\infty) \sim -(i/\pi) \ln U$. The charges are redefinable by circuiting infinity, which induces $(n_m,n_e)\rightarrow (-n_m,-n_e - 2 n_m)$. Furthermore, changing the $\theta$ angle of $\tau$ by $2\pi$ simply switches the labelling of the monopole and dyon, leaving the two singular points $U_i$ physically equivalent, as the ${\bf Z_2}$ symmetry requires. For the pure gauge $N=2$ $SU(2)$ model, these boundary conditions correspond to an elliptic curve \begin{equation} \label{ellipticSU2} y^2 \ =\ (x - U_i)(x + U_i)(x - U) \ , \end{equation} where $U_2 = -U_1$ by the ${\bf Z_2}$ symmetry. The period integrals are \begin{eqnarray} \label{aadintegrals} A &\ =\ & {\sqrt{2}\over\pi} \int^{U_i}_{-U_i} {dx \sqrt{x - U} \over \sqrt{x^2-U_i^2}} \\ && \nonumber \\ A_D &\ =\ & {\sqrt{2}\over\pi} \int_{U_i}^U {dx \sqrt{x-U} \over \sqrt{x^2-U_i^2}}\ . \end{eqnarray} These integrals can be expanded in powers of $z \equiv (U - U_i)/U_i$ with appropriate hypergeometric function expansions \cite{abramowitz}, yielding near $U = U_i$ \begin{eqnarray} \label{zexpansion} A &\ \sim\ &\sqrt{U_i} \ \left( \ {4\over\pi}\ +\ {z\over 2 \pi} ( 1 - \ln{z\over 32} ) \ +\ {z^2\over 32 \pi} ( {3\over 2}\ +\ \ln{z\over 32} ) \ +\ {\cal O}(z^3) \right) \nonumber \\ A_D &\ \sim\ & i\sqrt{U_i} \ \left( {z\over 2} \ -\ {z^2\over 32} \ +\ {3 z^3\over 2^9}\ + \ {\cal O}(z^4) \right) \nonumber \\ \tau_D &\ \sim\ & {i\over\pi} \left( -\ln{z\over 32}\ +\ {z\over 4} \ -\ {13\over 256}z^2 \ +\ {\cal O}(z^3)\right)\ , \end{eqnarray} The singular point $U_i$ can be obtained by matching to the one-loop beta function result, in the perturbative large-$U$ regime, and up to threshold corrections represented by a constant $c$ of order one, is \begin{equation} \label{Lambdadefn} U_i \approx \Lambda^2 \equiv c\Lambda_{UV}^2 \exp{(i\pi\tau_0)}\ , \end{equation} where $\Lambda_{UV}^2$ is the Wilsonian scale associated with the ``bare'' Lagrangian. The scalar potential in the dual theory, in the vicinity of either singular point where the dual theory is weakly coupled, is thus (to order $|a_D|^2 / \Lambda^2$) \begin{equation} \label{scalarpotential} V\ \approx\ 2 |a_D|^2 \left(|a_M|^2 + |a_{\tilde M}|^2\right) \ -\ \pi^2 \ { \left(|a_M|^2 - |a_{\tilde M}|^2\right)^2 \ +\ 16 |a_M a_{\tilde M}|^2 \over 2\ln{|a_D/16\Lambda|} }\ . \end{equation} This is minimized by $a_M = a_{\tilde{M}} = 0$, leaving $a_D$ as a quantum moduli space. At $a_D = 0$ the monopoles are massless. Before introducing $N=2$ supersymmetry breaking mass terms into the model, it is worth discussing the region of validity of the effective theory discussed above. The Wilsonian scale $\mu^2$ must stay below the scale $U$ where the $SU(2)$ gauge group is broken to the $U(1)$ subgroup. Furthermore, the monopole fields must have masses in excess of $\mu$, since we have cut off the monopoles' contribution to the $\tau$ function at their mass $\sim (U-U_i)/\sqrt{U_i}$. (If instead the monopole mass were less than $\mu$, then some of the gauge coupling's running would be due to loops computed in the low energy effective theory with internal momenta below $\mu$, and the $\tau$ function would not depend on $(U-U_i)/\sqrt{U_i}$.) The vacuum structure of the theories can be obtained from the limit $\mu \rightarrow 0$, so the effective theory is useful even close to the point $a_D = 0$ where the monopoles are light. If we are to extend the $N=2$ model's effective theory to models with $N=2$ breaking masses these models must continue to exhibit an energy range with a $U(1)$ gauge symmetry. As the $N=2$ breaking masses grow we expect the theories' mass gap to increase (since the cancellations preventing confinement of the $SU(2)$ gluons will break down). For an energy range with a $U(1)$ gauge symmetry to exist, the scale at which the $U(1)$ gauge boson is confined by the $SU(2)$ gauge dynamics, and also therefore the $N=2$ breaking masses, must be less than the $SU(2)$ breaking scale $U$. The resulting tractable models occupy a ``ball'' in parameter space around the $N=2$ model. At tree level these models possess an $SU(2)$ gauge symmetry; a light adjoint of fermions; and a light adjoint of scalars, not necessarily degenerate with the fermions. We cannot rule out the possibility that the models' qualitative properties change dramatically as the soft breakings are taken larger than $\Lambda$ (necessary to recover the QCD limit), although for small soft breakings we find that the behavior is smooth. Even though introducing $N=2$ breaking terms lifts the degeneracy of the moduli space, we can still discuss monodromies of the effective $\tau$ function in what was the moduli space by coupling a chiral superfield source $J(x)$ to the adjoint field $\Phi(x)$. The additional term \begin{equation} \label{source} {\cal L}_J\ =\ \int d^2 \theta\ J \Phi \ +\ {\rm h.c.} \end{equation} shifts $\langle a_{\Phi} \rangle$ away from its ($J=0$) vacuum value when $\langle f_J \rangle \neq 0$. The additional interaction in (\ref{source}) can of course induce new $J$-dependent terms in the effective Lagrangian ${\cal L}_{eff}^{J}$. However, for small $N=2$ breaking masses a background source with magnitude of order the small breaking mass is sufficient to explore the monodromies around the original $N=2$ singular points. In the case of the gauge kinetic term, the lowest order term induced by $J$, \begin{equation} \label{Jinduced} (\ J^\dagger J \ (D W / \Lambda^3)^2 \ ) \Big\vert_D \ \sim\ \vert f_J / \Lambda^3 \vert^2 \ F_{\mu\nu}^2 \ , \end{equation} is clearly a subleading contribution near the $N=2$ limit, leaving the monodromies and singularities of the broken model still controlled by the $N=2$ original $\tau$ function. We therefore expect to find light monopoles or dyons in the effective theory for $\langle a_\Phi \rangle$ near the $N=2$ singular points. \subsection{Breaking to $N=1$ supersymmetry} \label{sec-2.3} The $N=2$ supersymmetric model discussed in section~\ref{sec-2.2} can be broken to an $N=1$ model by introducing a mass term for the adjoint matter fields \cite{SW,IS}, corresponding in the enlarged model to allowing $\langle a_m \rangle \neq 0$. The effective Lagrangian is thus that of the $N=2$ model plus terms, invariant under the $N=1$ supersymmetry and gauged and global $U(1)$ symmetries of the model, that vanish as the adjoint mass is turned off. The effective superpotential can therefore receive corrections of the form \begin{equation} \label{N=1F} \Delta W \ =\ m f(U/\Lambda, \tau_0)\Big|_F \end{equation} where $f$ is an unknown function. This unknown mass renormalization arises from interactions of $A$ with $SU(2)$ gauge bosons at scales above $U$. Introducing this mass term can also give rise to additional $D$-terms of the form \begin{equation} \label{N=1D} \Delta {\cal L}_D \ =\ m^\dagger m\ G \left( {A_D \over \Lambda} , {A_D^\dagger \over \Lambda}, \tau_0, \tau_0^\dagger, {MM^\dagger \over \Lambda^2}, {\tilde{M} M^\dagger \over \Lambda^2}, {\tilde{M} \tilde{M}^\dagger \over \Lambda^2} \right)\Bigg|_D \end{equation} plus higher dimension terms in the gauge fields. In general $G$ is some complicated unknown function\footnote{ Functions like $G$ induced by supersymmetry breaking appear in $D$-terms of the effective Lagrangian, here and below. While there are some constraints on these functions, for example from weak coupling behavior and nonsingularity of the Kahler metric, these are usually not sufficient to globally determine the behavior.}. Note that the corrections in (\ref{N=1D}) to the kinetic energy terms are suppressed by ${\cal O} (m^2 / \Lambda^2)$,and for small $m$ we can consider them ${\cal O} (p^4)$. In this sense these corrections do not destroy the ${\cal O}(p^2)$ exactness of the $N=2$ solution. Upon eliminating $f_{A_D}$ from such terms and the canonical $D$-term for $A_D$, the potential only receives contributions of the form \begin{equation} \label{VfromWG} {1\over 1 + G^\prime\big|_A}\left| {dW\over df_{A_D}}\right|^2 \end{equation} with $G^\prime$ a function related to $G$. Now, supersymmetric vacua satisfy $V = 0$, and $G$ is nonsingular unless the effective theory completely breaks down. The extrema of $W$ thus coincide with the minimum of the potential, and we have \begin{equation} \label{neq1eqmot} \begin{array}{c} \sqrt{2}\, a_M a_{\tilde M}\ +\ m \left( d a_U / d a_D \right)\ =\ 0 \ \ \\ \\ a_D a_M\ =\ a_D a_{\tilde M}\ =\ 0 \ . \end{array} \end{equation} As Seiberg and Witten found, the potential is minimized for $\langle a_D \rangle\ =\ 0$ and the monopoles condense with \begin{equation} \label{neq1minimum} a_M\ =\ a_{\tilde M}\ =\ \left( -a_m {a_U}'(0) / \sqrt{2} \right)^{1/2}\ . \end{equation} The theory possesses a mass gap, since the magnetic $U(1)$ gauge boson acquires a mass by the Higgs mechanism, and electric charges are confined. \subsection{Breaking to $N=0$ with nonzero $f_m$} \label{sec-2.4} As a first example of soft supersymmetry breaking consider giving a nonzero vacuum expectation value to the $f$-component of $m$ in the $N=1$ model. At tree level this generates an extra $N=1$ breaking mass term for $a_\Phi$ \begin{equation} \label{fmterm} 2 {\rm Re}( f_m {\rm Tr}(a_\Phi^2) ) \ , \end{equation} which is renormalized by gauge interactions, through the function $f$ in (\ref{N=1F}), but remains in the effective theory. In addition the $D$-terms in (\ref{N=1D}) may generate additional terms \begin{equation} \label{evalfmterm} |f_m|^2 \ G\left({A_D \over \Lambda}, {A_D^\dagger \over \Lambda}, \tau_0, \tau_0^{\dagger}, {MM^\dagger \over \Lambda^2}, {\tilde{M} M^{\dagger} \over \Lambda^2}, {\tilde{M} \tilde{M}^\dagger \over \Lambda^2} \right)\Bigg|_A \ +\ a_m f_m G \Big|_F \ +\ {\rm h.c.} \end{equation} Firstly, we note that the unconstrained terms are small (${\cal O}(p^2)$ or higher) when $m \ll \Lambda$ and $f_m \ll \Lambda^2$. Secondly, introducing the soft breaking mass leaves the lowest dimension gauge kinetic term unchanged from the $N=2$ model (ignoring not only the higher dimension terms mentioned above, but also suppressed terms such as $m^\dagger m(DW)^2/\Lambda^4|_D \sim |f_m^2/\Lambda^4|(F_{\mu\nu})^2$). We conclude from the singularity structure of $\tau$ that even in the softly broken model there remain two points at which monopole bound states become massless. The soft breaking may generate $a_D$ interactions of unknown sign and similarly contribute to the masses of the scalar components of the monopole fields. When $f_m$ is the only $N=2$ breaking parameter, these terms are suppressed by $f_m/\Lambda$ relative to the bare scalar mass term, but unfortunately the tree level potential is then unbounded below as $\langle a_U \rangle \rightarrow -\infty$. Higher dimension $D$-terms, contributing for example $|a_U|$ terms to the potential, determine whether it is truly unbounded or is instead minimized at some finite $a_U$. Such unknown terms may however be small compared to an additional $N=2$ breaking mass generated by $\langle a_m \rangle \neq 0$. In the limit $f_m/\Lambda^2 \ll a_m/\Lambda \ll 1$, the bare masses dominate the corrections to the pure $N=2$ model and we obtain the potential \begin{eqnarray} \label{fmscalarpotential} V\ \approx\ {-8 \pi^2\over \ln{|a_D/16\Lambda|} } \left| a_M a_{\tilde{M}} - {i \sqrt{2} \Lambda a_m }\right|^2 \ +\ { 2 |a_D|^2 } \left(|a_M|^2 + |a_{\tilde{M}}|^2\right) \nonumber\\ \ -\ {\pi^2\over 2 \ln{|a_D/16\Lambda|}} \left(|a_M|^2 - |a_{\tilde{M}}|^2\right)^2 \ +\ 4 {\rm Im} (f_m \Lambda a_D) \end{eqnarray} which is minimized (via (\ref{neq1minimum}), up to terms of order $f_m^2$ and $a_m^2$) by $a_M = a_{\tilde M} = (i \sqrt{2} \Lambda a_m )^{1/2}$ and $a_D = i 2 \sqrt{2}f_m^*/a_m$. We thus obtain an ${\cal O} (p^0)$ solution of a model with $N=0$ supersymmetry. We conclude that up to small corrections the vacuum structure of a model with both a small $N=1$ preserving mass and a small soft supersymmetry breaking mass is equivalent to that of the pure $N=1$ model. There thus exist $N=0$ models which are close in parameter space to Seiberg and Witten's $N=1$ model that exhibit monopole condensation and confinement. Unfortunately the terms induced in the potential by soft breaking from $D$-terms are not determined by the super-, gauge or global symmetries of the model and thus we draw no conclusions as to the vacuum structure when $\langle a_m \rangle \rightarrow 0$. This theme will recur in the discussion below. \subsection{Breaking to $N=0$ with a gaugino mass} \label{sec-2.5} The $N=2$ model can be perturbed directly to an $N=0$, softly broken supersymmetric model by including a gaugino mass term. In the enlarged model such a mass can arise from setting $\beta_\tau \neq 0$ and hence $\langle f_{\tau_0} \rangle \neq 0$. The effective Lagrangian is again that of the $N=2$ model plus, potentially, all additional terms consistent with the symmetries of the model that vanish as the gaugino mass is switched off. The lowest dimension gauge kinetic term is given by the $\tau$ function of $N=2$ and hence the singularities indicating the presence of massless monopole fields are unchanged by the soft supersymmetry breaking. There are however extra terms generated in the dual theory close to $U_i$, arising from the terms in the $N=2$ effective Lagrangian when $\langle f_{\tau_0} \rangle \neq 0$. In particular, $f_{\tau_0}$ enters through $U_i$'s dependence (\ref{Lambdadefn}) on $\tau_0$; to second order in $a_z$, with $z \equiv \left(U - U_i \right)/U_i$, we have \begin{eqnarray} \label{dualz2potential} V &\approx & \left|{\Lambda f_{\tau_0} \over 16}\right|^2 \left( -\ln{|a_z|\over 32} + {a_z + a_z^*\over 8}\left(\ln{|a_z|\over 32} + 1\right) \right)^{-1} \nonumber\\ && \left| 8 - {32\sqrt{2}\pi \over \Lambda f_{\tau_0}} a_M^* a_{\tilde M}^* - 2 a_z\left(\ln{|a_z|\over 32} - 1/2\right) - a_z^*\left( 1 - {4\sqrt{2}\pi\over f_{\tau_0} \Lambda} a_M^* a_{\tilde M}^*\right) \right|^2 \nonumber\\ && - \left|{\Lambda f_{\tau_0} \over 4}\right|^2 \left(a_ - {a_z^2 \over 16} - {4\sqrt{2}\pi \over \Lambda^* f_{\tau_0}^*} a_M a_{\tilde M} a_z \right) + {\rm h.c.} \nonumber\\ && + {\left|\Lambda\right|^2 \over 2}|a_z|^2 \left(|a_M|^2 + |a_{\tilde M}|^2\right) \nonumber\\ && + {\pi^2 \over 2} \left(-\ln{|a_z|\over 32} + {a_z + a_z^*\over 8} \right)^{-1} \left(|a_M|^2 - |a_{\tilde M}|^2 \right)^2 \ . \end{eqnarray} There may in addition be $D$-terms that vanish in the $N=2$ limit and give contributions to the superpotential of the form \begin{equation} \label{ftaudirectD} \Lambda^2 H(\tau_0^\dagger, \tau_0, z, z^\dagger)\Big|_D \ . \end{equation} Such terms may induce a scalar monopole mass contribution, for example \begin{equation} \label{mpolemassexample} U_i^\dagger z \Big|_D \sim\ \Lambda^2 f_{\tau_0}^* f_z \ , \end{equation} which on eliminating $f_z$ provide a shift proportional to $f_{\tau_0}$ within $|f_z|^2$ in (\ref{dualz2potential}). The resulting cross terms with the monopole fields induce a scalar monopole mass. Its sign and magnitude, which would signify whether they condense, is therefore undetermined by the symmetries. These terms perturb, in an unknown fashion, the potential (\ref{dualz2potential}). One can however read off the gaugino mass in the effective theory, arising from the $f$-component of the $\tau$ function, as \begin{equation} \label{gluinomass} m_{\tilde{\gamma}_D} \ =\ {m_{\tilde\gamma} \over a_{\tau}} \ \sim\ \tau_D^\prime (a_{\tau_0}) f_{\tau_0} \ \sim\ -f_{\tau_0} {\Lambda^2 \over a_D}\ . \end{equation} The gaugino and dual gaugino masses diverge as $a_D \rightarrow 0$, which we may interpret as the decoupling of the massive gaugino fields from the effective theory, as the decreasing monopole mass drives its region of validity to zero. If the theory behaves without singularity as $\langle f_{\tau_0} \rangle \rightarrow 0$, the contribution of (\ref{ftaudirectD}) to the potential must cause the vacuum expectation $\langle a_D \rangle$ to become nonzero and proportional to $\langle f_{\tau_0}\rangle ^l$ with $l<1$, so that $m_{\tilde{\gamma}_D}$ smoothly approaches zero in the $N=2$ limit. As before if the gaugino mass is introduced as a perturbation to the $N=1$ model ($\langle f_{\tau_0} \rangle \ll \langle a_m \rangle$), then the corrections to the potential from the gaugino mass will be small and the minima will be that of the $N=1$ model with monopole condensation. \subsection{Breaking to $N=0$ with a squark mass} \label{sec-2.6} Allowing $\langle f_K \rangle \neq 0$ in (\ref{smallLagrangian}) generates a soft supersymmetry breaking squark mass, induced at tree level by the term $|f_K|^2 |a_{\Phi}|^2$. Since $K$ has arbitrary $U(1)$ charge it could not appear in the $F$-terms of the effective theory, but it does appear in $D$-terms of the form \begin{eqnarray} \label{effectiveKterm} {\cal L}_K &\ =\ &\left[ K^\dagger K\ \Phi^\dagger \Phi\ G_1 \left(\tau_0, \tau_0^\dagger, {\Phi^\dagger \Phi \over \Lambda^2}, {\Phi^2 \over \Lambda^2}, {M^\dagger M \over \Lambda^2}, {{\tilde M}^\dagger {\tilde M} \over \Lambda^2} \right) \right. \nonumber \\ && \left. \ +\ \ K^\dagger K\ \Phi^2 \ G_2 \left(\tau_0, \tau_0^\dagger, {\Phi^\dagger \Phi \over \Lambda^2}, {\Phi^2 \over \Lambda^2}, {M^\dagger M \over \Lambda^2}, {{\tilde M}^\dagger {\tilde M} \over \Lambda^2} \right) \right]_D \ , \end{eqnarray} with $G_1$ and $G_2$ unknown functions. The bare squark mass is preserved up to gauge renormalization, but the induced monopole mass is undetermined and hence so is the vacuum structure. The $\tau$ function of the $N=2$ model is again preserved, indicating the presence of massless monopole states at $a_D = 0$. The potential minima when the $N=1$ model is perturbed will still lie close (for small $\langle f_K \rangle$) to $a_D = 0$ and $a_M$ given by (\ref{neq1minimum}), not altering the vacuum structure from the pure $N=1$ case. \section{$SU(N_c)$ Gauge Symmetry And Matter Fields} \label{sec-matter} The exact results for $N=2$ supersymmetric QCD with an $SU(2)$ gauge group have been generalized to the case of an $SU(N_c)$ gauge group \cite{Ncolor}. The potential for the adjoint scalars is minimized by a vev that classically breaks $SU(N_c)$ to $U(1)^{N_c}$. For large scalar vev the $U(1)$ couplings can again be calculated in perturbation theory and shown to be consistent with the existence of $N$ points on the quantum moduli space at each of which $N-1$ dyons become massless. The function $\tau$ is again exactly specified (up to scaling) by the singularity structure. Seiberg and Witten have also demonstrated how $N=2$ matter multiplets can be included in the $SU(2)$ model. For heavy matter fields, the $SU(2)$ dynamics are invariant at low energies, except for the addition of an extra singularity at large $U = m/\sqrt{2}$ corresponding to where a matter field mass vanishes due to its Yukawa coupling to the scalar vev. As the mass decreases, the singularity moves towards the origin of the moduli space and merges with the pure glue singularities as dictated by the requirement that the dyons fill out multiplets of the relevant flavor symmetries. Again the $\tau$ function is exactly determined by the monodromy structure. These results can also be generalized to $SU(N_c)$ gauge models \cite{matterN}. In each of these cases the exact form of the $\tau$ function in the effective theory is known (up to higher dimensional contributions). We can break the models to $N=0$ models with squark and gaugino masses by promoting the bare couplings $\tau_0$ and $K$ to the status of chiral superfields and allowing them to acquire nonzero $f$-component vevs. As for the $SU(2)$ theory, the $D$-terms will generate unknown contributions to the potential, but the gauge multiplet's kinetic term is given exactly by $\tau$. If the soft breaking masses are small relative to an $N=1$ preserving adjoint matter mass then the models are pinned near the singularities and the monopoles condense. Among these models is the particularly interesting case of an $SU(3)$ gauge symmetry and $N=1$ matter multiplets in the fundamental representation. We may choose to introduce an $N=1$ supersymmetry preserving mass for the adjoint matter field and soft supersymmetry breaking masses for the gauginos and fundamental representation scalars. At tree level these masses can be raised until the fields are integrated from the theory, leaving QCD with quarks. Although we cannot explicitly take this limit in the effective theory, since the light degrees of freedom there are not those appropriate to QCD, we may begin the interpolation from the $N=2$ result towards QCD. The behavior of the theory is that monopoles condense and the quarks are confined. This picture's consistency with the t'Hooft--Mandlestam picture \cite{dualqcd} of quark confinement in QCD suggests that the models are smoothly connected. \section{Discussion} \label{sec-discussion} We have embedded the $N=2$ and $N=1$ supersymmetric models possessing exactly calculable results within a larger $N=1$ model, which in certain limits induces soft supersymmetry-breaking masses in the original theories. This procedure generates new contributions to the effective Lagrangian, which unfortunately holomorphy and $U(1)$ symmetries do not completely determine. Thus the information we are able to extract about the behavior of corresponding $N=0$ models is limited. However, at leading order in the breaking masses the gauge kinetic terms are still completely determined, leaving unchanged the $\tau$ function of the $N=2$ and $N=1$ models. Since the singular behavior of $\tau$ is consistent with the original ansatz that a monopole becomes massless at each of the singular points, we expect that these are still the only singularities in the low-energy effective Lagrangian of the nonsupersymmetric models. If the soft supersymmetry breaking masses are induced as sufficiently small perturbations to the $N=1$ preserving model, then monopole condensation and confinement are preserved and seen not to depend on the presence of exactly massless gauginos or squarks, nor on exact supersymmetry degeneracies. While we have encountered some limitations in applying Seiberg and Witten's techniques to nonsupersymmetric models, the analysis nevertheless reveals some details of condensation and confinement. Our results are consistent with the supersymmetric models being smoothly connected to nonsupersymmetric QCD-like models. \begin{flushleft} {\Large\bf Acknowledgments} \end{flushleft} The authors are grateful to Luis Alvarez-Gaum\'e, Jacques Distler, Mans Henningson, Greg Moore, Nati Seiberg and Matt Strassler for useful discussions and comments. This work was supported in part under United States Department of Energy contract DE-AC02-ERU3075. SDH acknowledges the hospitality of the Aspen and the Benasque Centers for Physics, where some of this work was performed. \newpage \baselineskip=1.6pt
\section{Introduction} The COBE experiment (Bennet et al. 1992, 1994; Smoot et al. 1992) has stimulated a considerable amount of work on cosmic structure. Current tests usually exploit the angular correlation function and several harmonic amplitudes of the sky temperature field (see e.g, Adams et al. 1992; Kashlinsky 1992; Efstathiou et al. 1992; Kofmann et al. 1993; Gorski et al. 1994). However, several more tests have been suggested over the years for a thorough investigation of the properties of the anisotropy field of the cosmic background radiation. These often involve the distribution and features of hot and cold spots, which can provide useful checks of the Gaussian nature of the fluctuations (Sazhin 1985; Bond \& Efstathiou 1987; Coles \& Barrow 1987; Coles 1988; Mart\'\i nez-Gonz\'alez \& Sanz 1989; Gott et al. 1990). Measurable quantities include the number of spots $N_{{\rm iso}% }$ defined by isotemperature contours, the spot boundary curvature or genus $% G$, the spot excursion area, and so on. A first analysis of COBE-DMR maps along these lines has been performed by Torres (1994). A definition of spots independent of isotemperature contours considers local maxima and minima of temperature (Bond \& Efstathiou 1987, Vittorio \& Juszkiewicz 1987), and is thereby not connected to the topological features of spots. Differing from $N_{{\rm iso}},$ the dependence of the number of (positive and/or negative) peaks $N_{{\rm peak}}$ on threshold is not universal for Gaussian fields. This latter approach was adopted by Fabbri \& Natale (1993, 1995) in studies of the 2-dimensional distribution of extragalactic IRAS sources, but has not yet been applied to the cosmic background radiation. In this work we analyze the statistics of local maxima and minima in COBE-DMR 2-year maps. We found that in this kind of analysis the detector noise must be taken into account very carefully since $N_{{\rm peak}}$ is sensitive also to high order harmonics where noise dominates (cf. Fabbri 1992). However, due to a highly nonlinear dependence of $N_{{\rm peak}}$ on the harmonic strengths, the presence of structured signals in COBE maps reduces its value below the level measured in pure noise maps. (The identification of genuine peaks in the radiation temperature is not required at all in our analysis.) We find that the distributions of positive and negative peaks are mutually consistent, and the results from this statistics agree with those of earlier tests. Therefore, we find no evidence for non-Gaussian features in the fluctuations. More precisely, fitting Gaussian power-law models of cosmic structure to the peak distribution we recover a clear anticorrelation between the spectral index $n$ and the predicted rms quadrupole $Q_{{\rm rms-PS}}$ (Seljak \& Bertschinger 1993; Smoot et al. 1994): We get $Q_{{\rm rms-PS}}=17\pm 3$ $\mu $K for $n=1$ and $14\pm 3$ $% \mu $K for $n=1.5$, where the error bars include uncertainties deriving from the treament of noise as well as cosmic variance. These numbers, altough they agree with previous evaluations of the quadrupole from higher order harmonics, are not consistent with its direct determination providing $Q_{% {\rm rms}}=6\pm 3$ $\mu $K (Bennet et al. 1994). So the recently discovered discrepancy is confirmed by the properties of the peak distribution, which depend on the harmonic content of the angular distribution up to $\ell \sim 50.$ \section{The peak number test} We analyzed the 2-year 53(A+B) DMR maps processed with a 2.9$^{\circ }$ smoothing (Wright et al. 1993) and dipole subtraction. Considering the Northern and Southern hemispheres separately, we constructed two pole-centered maps, each containing 12892 pixels, using the coordinate transformation $\theta _1=2\sin \left[ \frac 12\left( \frac \pi 2-\left| b\right| \right) \right] ,$ $\phi _1=l,$ with $b$ and $l$ the Galactic coordinates. After masking low Galactic latitudes, $\left| b\right| <20^{\circ },$ we were left with 8412 pixels per map. We then looked for temperature peaks using the algorithms of Fabbri \& Natale (1993, 1995). Table 1 gives the no-threshold numbers of peaks, both actually detected and extrapolated to the entire sky and to the North and South hemispheres (2nd and 3rd column, respectively). Figure 1 reports the extrapolated numbers vs. a threshold factor $\nu $. This is the peak height normalized to the sky rms fluctuation $C^{\frac 12}(0)$; for cold spots, $N_{{\rm peak}}$ gives the number of minima below $-C^{\frac 12}(0)\nu $. For distributions of only positive {\it or} negative peaks the statistical errors at 1-sigma confidence levels are evaluated as $\left( N_{{\rm peak}}/f_{{\rm U}}\right) ^{\frac 12},$ with\/ $f_{{\rm U}}=0.652$ the unmasked fraction of the sky. Within such error limits, we find no significant difference between the distributions of positive and negative peaks; this result provides support for the Gaussian nature of cosmic perturbations. In the figure we also report the full-sky average number of positive and negative peaks. This average will be compared with theoretical models below, because of the smaller (by a factor $\sqrt{2})$ relative error. \begin{table*} \caption[ ]{No-threshold peak numbers \label{peaks}} \begin{flushleft} \begin{tabular}{llll} \hline \noalign {\smallskip} Peak set & $(A+B)$ Maps & $(A+B)$ Maps & $(A-B)$ Maps\\ & $|b| > 20^{\circ}$ & Extrapolated$^{\dagger }$ & No mask \\ \noalign {\smallskip} \hline \noalign {\smallskip} Hot, North & 50 & $76.6 \pm 10.8 $ & 95 \\ Cold, North & 47 & $72.0\pm 10.5 $ & 99 \\ Hot, South & 48 & $73.6 \pm 10.6 $ & 105 \\ Cold, South & 53 & $81.2 \pm 11.2 $ & 97 \\ Hot, total & 98 & $ 150.2 \pm 15.2 $ & 200 \\ Cold, total & 100 & $ 153.3 \pm 15.3 $ & 196 \\ Hot/Cold Average & 99 & $ 151.7 \pm 10.8 $ & 198 \\ \noalign {\smallskip} \hline \end{tabular} $^{\dagger }$To full sky or hemispheres, including masked regions \end {flushleft} \end{table*} \begin{figure} \picplace{8.0cm} \caption[ ] {The peak number $ N_{\rm peak}$ vs. threshold in COBE-DMR maps. Open squares (triangles) denote positive (negative) peaks. The four lowest curves give the peak numbers extrapolated to the Northern (dotted lines) and Southern (dash-dotted) halves of the sky. The remaining curves refer to the full sky. Filled circles give the average numbers of positive and negative peaks. The full line describes a power-law model with $n = 1$ and $Q_{% {\rm rms-PS}} = 18.5$ $ \mu $K, corresponding to the fit procedure S31 in Table 2} \end{figure} \begin{figure} \picplace{8.0cm} \caption[ ] {The noise harmonic coefficients $A_{{\rm N}\ell }^2$ (containing the form factors $W_\ell ^2$) and the beam shape coefficients $G_\ell $ vs. $\ell.$ Filled circles give $A_{{\rm N}\ell }^2$ as obtained from the noise map, and the dotted line represents the best fit curve corresponding to $\sigma _{\rm N}=0.936^{\circ }$ and $C_{\rm N}=5.74\times 10^{-5}$ mK$^2.$ Crosses give the experimental $G_\ell $, compared to Gaussian approximations with dispersion 2.9$^{\circ }$ (full line) and 3.1$^{\circ }$ (dash-dotted) } \label{figure1} \end{figure} For the interpretation of the above data, we must consider that both cosmic signal and noise contribute to the temperature field $T(\theta ,\phi )=\sum a_{\ell m}Y_{\ell m}(\theta ,\phi )$. The expectation values of the numbers of local maxima and minima are determined by the variances $a_\ell ^2=\sum_m\left\langle \mid a_{\ell m}\mid ^2\right\rangle $. For a Gaussian field the full-sky $N_{{\rm peak}}$ is given by \begin{eqnarray} \label{Npeak} & &\left\langle N_{{\rm peak}}(\nu )\right\rangle = \sqrt{\frac 2\pi } \theta^{*}{}^{-2}\left\{ \gamma ^{*2}\nu {\rm exp}\left( -\frac 12\nu ^2\right) \right. \nonumber \\ & &+ \frac{\gamma ^{*}\left( 1-\gamma ^{*2}\right) ^{\frac 32}}{\sqrt{2\pi }} {\rm % exp}\left( -u_2^2\nu ^2\right) +\sqrt{\frac \pi 6}{\rm erfc}\left( \sqrt{% \frac 32}u_1\nu \right) \nonumber \\ & & - \frac 12\int_\nu ^\infty \left[ \gamma ^{*2}(y^2-1){\rm % exp}\left( -\frac 12y^2\right) {\rm erfc}\left( \gamma ^{*}u_2y\right) \right. \nonumber \\ & &+ \left. \left. u_1% {\rm exp}\left( -\frac 32u_1^2y^2\right) {\rm erfc}\left( \gamma ^{*}u_1u_2y\right) \right] {\rm d}y \right\} , \end{eqnarray} where \begin{equation} \label{u1u2}u_1=\left( 3-2\gamma ^{*2}\right) ^{-\frac 12},\;\;\;u_2=\left[ 2\left( 1-\gamma ^{*2}\right) \right] ^{-\frac 12}, \end{equation} and the properties of the anisotropy field are summarized by the parameters $% \theta ^{*}$ and $\gamma ^{*}$ (Bond \& Efstathiou 1987; Fabbri 1992) given by \begin{eqnarray} \label{tetastar} \theta ^{*2} & = & \left[ 2\sum_\ell \ell \left( \ell +1\right) W_\ell ^2a_\ell ^2\right] \nonumber \\ & \times & \left[ \sum_\ell \left( \ell -1\right) \ell \left( \ell +1\right) \left( \ell +2\right) W_\ell ^2a_\ell ^2\right] ^{-1}, \end{eqnarray} and \begin{eqnarray} \label{gammastar}\gamma ^{*2} & = & \left[ \sum_\ell \ell \left( \ell +1\right) W_\ell ^2a_\ell ^2\right] ^2\left[ \sum_\ell W_\ell ^2a_\ell ^2\right]^{-1} \nonumber \\ & \times & \left[ \sum_\ell \left( \ell -1\right) \ell \left( \ell +1\right) \left( \ell +2\right) W_\ell ^2a_\ell ^2\right] ^{-1}. \end{eqnarray} Here $W_\ell $ are form factors taking into account beam shape and any additional smearing effect. Assuming that signal and noise are uncorrelated, their contributions to $% a_\ell ^2$ add up in quadrature in the (A+B) maps, $a_\ell ^2=a_{{\rm S}\ell }^2+a_{{\rm N}\ell }^2$, and we need to determine $a_{{\rm N}\ell }^2$ in an independent way. This can be achieved by means of the (A$-$B) maps, which can however be used to directly derive the coefficients $A_{{\rm N}\ell }^2=W_\ell ^2a_{{\rm N}\ell }^2$ rather than the $a_{{\rm N}\ell }^2$. Figure 2 reports the results of a harmonic best fit up to $\ell =30$ executed on the entire celestial sphere$.$ Increasing the number of harmonics, best fits (as well as direct integration by means of $a_{\ell m}=\int T(\theta ,\phi )Y_{\ell m}^{*}(\theta ,\phi ){\rm d}\Omega $) tend to overrate large-$\ell $ amplitudes. We checked for this effect by considering the peak statistics for noise maps.% \begin{figure} \picplace{8.0cm} \caption[ ] {The peak number $ N_{\rm peak}$ vs. threshold in COBE-DMR noise maps. Open squares (triangles) denote positive (negative) peaks. The four lowest curves give the peak numbers extrapolated to the Northern (dotted lines) and Southern (dash-dotted) halves of the sky. The remaining curves refer to the full sky. Filled circles give the average statistics of positive and negative peaks. The full line represents the best fitting function} \label{figure3} \end{figure} \begin{figure} \picplace{6.5cm} \caption[ ] {The acceptance regions in the $(n,Q_{{\rm rms-PS}})$ plane at 1- and 2-sigma confidence levels. Contours are computed with procedures S31 (full lines) and F31 (dotted). Both procedures assume $\sigma_{\rm S} = 3.1^{\circ}$, $\sigma_{\rm N} = 0.936^{\circ},$ and $C_{\rm N} =5.74 \cdot 10^{-5} $ mK$^2$. The former uses the measured $A_{\rm N\ell }^2$ up to $ \ell = 30$.} \label{figure4} \end{figure} Figure 3 gives the numbers of maxima and minima detected in the full-sky noise maps vs. threshold $\nu .$ If using Eq.s (\ref{Npeak}-\ref{gammastar}) we generate the peak statistics from a set of 50 harmonics with best-fitted amplitudes, we get a low-threshold excess of about 30 peaks with respect to the data in Fig. 2; the discrepancy increases with the number of harmonics. Since accurate theoretical calculations of peak statistics require at least $% \sim $50 harmonics, we tried to fit an analytic form to a more limited set of $A_{{\rm N}\ell }^2$. A satisfactory choice is \begin{equation} \label{anoise}A_{{\rm N}\ell }^2=C_{{\rm N}}\left( \ell +\frac 12\right) \exp \left[ -\left( \ell +\frac 12\right) ^2\sigma _{{\rm N}}^2\right] . \end{equation} However the large error bars on individual $A_{{\rm N}\ell }^2$ make a 2-parameter fit for the function (\ref{anoise}) not very useful. A more convenient solution to this problem is to fit the parameter $\sigma _{{\rm N}% }$ directly on the noise-generated peak distribution; using the average distribution of maxima and minima (represented by the filled circles in Figure 3) and 99 harmonics the result is $\sigma _{{\rm N}}=0.936^{\circ }\pm 0.071^{\circ }.$ (We checked however that a set of $\sim 50\div 60$ harmonics would provide a sufficient accuracy.) The fit turns out to be even ``too'' good, providing $\chi _{\min }^2=1.3$ against 9 degrees of freedom: This means that data points at different thresholds are not uncorrelated. The value found for $\sigma _{{\rm N}}$ is quite independent of $C_{{\rm N}}$% , since peak statistics does not depend on the overall amplitude of anisotropies as shown by Eq.s (\ref{tetastar}, \ref{gammastar}).We then fitted the amplitude parameter of Eq. (\ref{anoise}) on the reported set of $% A_{{\rm N}\ell }^2$ with $\ell =2\div 30,$ getting $C_{{\rm N}}=(5.74\pm 0.66)\times 10^{-5}$ mK$^2.$ The distribution described bu the full line in Fig. 3 was generated using the function (\ref{anoise}) in the entire range $% \ell =2\div 99.$ We checked that the peak distribution of the (A-B) maps is also recovered to a good accuracy building up the harmonic spectrum with the ``measured'' $A_{{\rm N}\ell }^2$ up to $\ell =30$ and the function (\ref {anoise}) at higher $\ell .$ We tested power-law spectra against the peak statistics in Fig. 1 with the following procedure. For each spectrum, labelled by the spectral index $n$ and the predicted quadrupole $Q_{{\rm rms-PS}}=a_{\rm S2}/\sqrt{4\pi },$ we generated the theoretical expectation values $a_{{\rm S}\ell }^2$ and then \begin{equation} \label{totalAl}W_\ell ^2a_\ell ^2=G_\ell ^2\exp \left[ -\left( \ell +\frac 12\right) ^2\sigma _{{\rm S}}^2\right] a_{{\rm S}\ell }^2+A_{{\rm N}\ell }^2. \end{equation} Here $G_\ell $ denote the measured beam-shape coefficients of COBE-DMR. These have reported by Wright et al. (1994) up to $\ell =50.$ A more extensive set of $G_\ell $ up to $\ell =99$ has been provided to us by Kogut (private communication). The exponential factor in Eq. (\ref{totalAl}) takes into account the 2.9$^{\circ }$ smearing on the map as well as the additional smearing due to orbital motion, so that our best estimate is $% \sigma _{{\rm S}}=3.1^{\circ }$. Note that $\sigma _{{\rm S}}$ should not be confused with the approximate Gaussian beamwidth which was used in many computations, but not in the present work. Also, it is quite different from the phenomenological parameter $\sigma _{{\rm N}}$ of Eq. (\ref{anoise}). The peak distributions calculated from Eq.s (\ref{Npeak}-\ref{gammastar}) are tested against the averaged distribution of positive and negative peaks. We should notice that such theoretical distribution are affected by several sources of uncertainty. Cosmic variance (White et al. 1993) affects the cosmic-structure contribution to the harmonic coefficients $a_{\ell m}$, and a quite similar variance pertains to noise; in fact, these two effects are described by identical equations assuming that both are Gaussian processes. Performing a limited number of simulations for superpositions of Gaussian signal and noise (with fixed $a_\ell ^2$, the {\em expectation} values of harmonic strengths), we found that the probability distribution of $N_{{\rm % peak}}$ at a given threshold is roughly consistent with a Poisson distribution having a width $\left\langle N_{{\rm peak}}\right\rangle ^{\frac 12}.$ However, in our case the $A_{{\rm N}\ell }^2$ themselves are not fixed. When we use Eq. (\ref{anoise}) their errors at different $\ell $ are correlated due to the uncertainties on $C_{{\rm N}}$ and $\sigma _{{\rm N% }};$ it can be shown that the corresponding contribution to the uncertainty on the predicted $N_{{\rm peak}}$ is given by \begin{eqnarray} \label{vars} & &{\rm var}(N_{{\rm peak}}) = \left( \sum_\ell A_{{\rm N}\ell }^2% \frac{\partial \left\langle N_{{\rm peak}}\right\rangle }{\partial A_{{\rm N}% \ell }^2}\right) ^2\frac{{\rm var}(C_{{\rm N}})}{C_{{\rm N}}^2} \nonumber \\ & & +4 \left[ \sum_\ell \left( \ell +\frac 12\right) ^2A_{{\rm N}\ell }^2\frac{\partial \left\langle N_{{\rm peak}}\right\rangle }{\partial A_{{\rm N}\ell }^2}% \right] ^2\sigma _{{\rm N}}^2{\rm var}(\sigma _{{\rm N}}). \end{eqnarray} When we use the measured $A_{{\rm N}\ell }^2,$ a more familiar equation $% {\rm var}(N_{{\rm peak}})=\sum_\ell \left[ \left( \partial \left\langle N_{% {\rm peak}}\right\rangle /\partial A_{{\rm N}\ell }^2\right) ^2{\rm var}(A_{% {\rm N}\ell }^2)\right] $ applies. For each computed model we calculated the corresponding $\chi ^2$ combining quadratically the above errors with with the experimental error bars reported in Fig.1. This procedure allowed us to avoid a more extensive use of simulations. Figure 4 gives the allowed regions at 1 and 2 sigmas in the $(n,Q_{{\rm % rms-PS}})$ plane. Full lines give the contours calculated for $\sigma _{{\rm % S}}=3.1^{\circ },$ using the $A_{{\rm N}\ell }^2$ of Fig. 2 up to $\ell =30$% , and the analytic form (\ref{anoise}) with the optimal values of $\sigma _{% {\rm N}}$ and $C_{{\rm N}}$ at higher $\ell .$ (This procedure is referred to as S31 in the Figure and in Table 2.) Contours calculated for $\sigma _{% {\rm S}}=2.9^{\circ }$ (case S29, not reported in the Figure) would be hardly distinguishable from them. Using the analytic form (\ref{anoise}) in the entire range $\ell =1\div 99$ (case F31) the contours are slightly displaced to smaller values of $Q_{{\rm rms-PS}}$ (by $\sim 2\ \mu $K for $% n=1)$. \begin{table} \caption[ ]{Coefficients of the $n-Q_{\rm rms-PS}$ regression defined by Eq. (\ref{ab}) \label{regress}} \begin{flushleft} \begin{tabular}{lll} \hline \noalign {\smallskip} Procedure & $a$ ($\mu$K) & $b$ ($ \mu$K) \\ \noalign {\smallskip} \hline \noalign {\smallskip} S31$^\dagger$ & $24.71\pm 0.65$ & $ -6.23\pm 0.45$ \\ S29 $^\ddagger$ & $24.54\pm 0.66$ & $ -6.16\pm 0.46$ \\ F31$^{\star }$ & $21.47\pm 0.60$ & $ -5.45\pm 0.42$ \\ \noalign {\smallskip} \hline \end{tabular} $^\dagger \sigma_{\rm N} =3.1^\circ$, using noise harmonic amplitudes from Fig. (2) for $\ell \leq 30$. $^{\ddagger }\sigma_{\rm N} =2.9^\circ$ $^{\star }$Noise described by Eq. (\ref{anoise}) for any $\ell $ \end {flushleft} \end{table} The parameters $n$ and $Q_{{\rm rms-PS}}$ are clearly anticorrelated, as already found from analyses of harmonic amplitudes and the angular correlation function (Seljak \& Bertschinger 1993; Torres et al. 1994; Smoot et a. 1994). Minimizing $\chi ^2$ for fixed $n$ we identify a straight line in the $(n,Q_{{\rm rms-PS}})$ plane, \begin{equation} \label{ab}Q_{{\rm rms-PS}}=a+bn, \end{equation} with the coefficient values listed in Table 2. From the above results, taking into account differences arising from the S31 and F31 procedures, we can conclude that $Q_{{\rm rms-PS}}=17\pm 3$ $\mu $K for $n=1$, and $Q_{{\rm % rms-PS}}=14\pm 3$ $\mu $K for $n=1.5$. These numbers agree very well with the most likely quadrupole $Q_{{\rm rms-PS}}$ derived from higher order multipoles, but not with the quadrupole rms fluctuation of 6$\pm 3$ $\mu $K directly fitted on two-year data; see Bennet et al. (1994) for a discussion of this discrepancy, and Jing \& Fang (1994) for a possible explanation in terms of an infrared cutoff in the spectrum. \begin{acknowledgements} We wish to thank A. Kogut for providing unpublished data on COBE-DMR beam shape. This work is partially supported by Agenzia Spaziale Italiana under Contract \# 94-RS-155, by the Italian Ministry for University and Scientific and Technological Research (Progetti Nazionali e di Rilevante Interesse per la Scienza), and by the European Union under Contract \# CI1*-CT92-0013. \end{acknowledgements}
\section{Introduction} One of the basic assumptions of the standard Friedmann cosmology is the Cosmological Principle which states that the Universe is homogeneous and isotropic\cite{Weinberg} so that every point of the Universe is equivalent. While isotropy has been reasonably well established from, among others, the observation of the cosmic microwave background radiation by COBE\cite{COBE}, homogeneity has been challenged by various observations of the Large Scale Structures (LSS). Recent galaxy redshift surveys such as CfA 1\cite{CfA1}, CfA 2\cite{CfA2}, SSRS1\cite{S1}, SSRS2\cite{CfA1} and the pencil beam surveys\cite{pencil} have provided evidence for the LSS such as filaments, sheets, superclusters and voids, up to $200-300 h^{-1}$Mpc. The current interpretation based on the observation of the LSS is that the Cosmological Principle may not be applicable, at least, to the local Universe, although it may be applicable to the Universe on a very large scale with a characteristic distance $\lambda$. However, one of the most remarkable consequences of the above galaxy surveys is that the scale of the largest structures in each survey is comparable with the extent of the survey itself. Recently, it has been suggested that from pencil beam surveys\cite{pencil} as well as from new deep redshift survey (ESP survey) \cite{ESP}, $\lambda$ should be much larger than the survey limits $\sim 500-600 h^{-1}$Mpc, implying the absence of any tendency towards homogeneity up to the present observational limits. For example, Pietronero\cite{fractal} has attempted to explain these phenomena by suggesting that the LSS shows fractal properties of the Universe. Even though this non--analytical distribution of matter means that the Universe is not homogeneous, the local isotropy preserves the fundamental assumption that every point of the Universe is equivalent. Unfortunately, however, a rigorous mathematical description of such a Universe is extremely difficult and in practice it is almost impossible. Therefore, it is desirable to simplify the description of this $inhomogeneous$ Universe to the extent that its analytical study becomes possible in order to see, at least, qualitative features of the matter distribution and cosmological consequences. History of cosmological models for an inhomogeneous Universe dates back to as early as 1930's. Lemaitre, Tolman and Dingle \cite{Lemaitre} attempted to describe the evolution of the $fluctuation$ in the mass distribution. Later in 1947, Bondi \cite{Bondi} elaborated the model and discussed observational consequences. In their model, which we shall call Tolman-Bondi(TB) model, the global Universe is that of the standard Friedmann cosmology, implying homogeneity over the region of order $\lambda$. Recently, in the frame work of the TB model, Moffat and Tatarski\cite{Moffat} studied, in order to describe the local inhomogeneity, cosmology of a local void in the globally Friedmann Universe and its effect on the measurement of the Hubble constant and the redshift--luminosity distance relation. Since the Universe modeled in \cite{Moffat} consists of many expanding voids and we happen to be located at the center of one of them, the shell--crossing singularity occurs, implying that different shells collide and the comoving coordinate become meaningless. Recently, based on the fact that there is no observational evidence of approaching towards the homogeneity within the survey limit, another cosmological model\cite{Kim} was proposed, whose global feature is not asymptotic to the Friedmann cosmology. In \cite{Kim}, the observable Universe is modeled as being inside an expanding bubble with the underdense center and matter inside the bubble is isotropically (but inhomogeneously) distributed when viewed by an observer located at the center. They proposed that such a Universe may be described by the following inhomogeneous metric \begin{equation} \di \tau^2 = \di t^2 - R^2(t,r)[\di r^2+r^2 \di \Omega^2], \end{equation} where $R(t,r)$ is the scale-factor, dependent on $r$ as well as on $t$. Therefore, homogeneity in the Cosmological Principle is explicitly violated, whereas isotropy remains intact. Based on the high degree of isotropy of the cosmic microwave background radiation measured by COBE\cite{COBE}, it was assumed that the observer is located at the center of the bubble (or near it), albeit the return of the pre--Copernican notion. Whether this picture is correct or not can only be decided when the results of the model are confronted with the observation. In \cite{Kim}, its cosmological consequences were qualitatively discussed. For example, the Hubble constant, the density parameter and the age of the Universe all became scale--dependent, whereas the analysis in \cite{Moffat} was simplified to avoid the possible position--dependent age. Moreover, because of the lack of the light propagation solution on which every observation is based, no explicit and testable cosmological results were derived, which can be compared with the observation. Therefore, it is interesting to examine observational consequences of such a model and compare with those of the TB model. In this paper, we first present a general redshift--luminosity distance relation for a certain class of inhomogeneous cosmological models. Then we apply the result to the model discussed in \cite{Kim} as an example. The plan of this paper is as follows. In Section II, we present the redshift--luminosity distance relation for the case of one $(t,r)$--dependent scale factor in the metric. In order to proceed further to derive some specific observable consequences, we have chosen the model proposed in \cite{Kim} as an example and briefly summarize the model in Section III. In Section IV, we derive modified results of the redshift--luminosity distance relation and show that they reduce to the well-known relations in the standard Friedmann cosmology for small $z$, i.e., for nearby objects. Section V deals with some numerical results of this model which are applicable for large $z$. Also discussed in this Section is the observed increase of density parameter with the redshift in the framework of this model. A brief summary and conclusions are given in Section VI. \section{Redshift in a General Metric with one Scale Factor } With the exception of astronomical neutrinos and possible future gravitational waves, most of the cosmological measurements are based on the electromagnetic waves, which travel along the null geodesic, i.e., $\di \tau=0$. Considering only the radial propagation ($\di \Omega =0$), we have, from Eq.(1), \begin{equation} \di r = - \frac{\di t}{R(t,r)}~, \end{equation} where the minus sign is chosen since $r$ decreases as $t$ increases. With a given $R(t,r)$, Eq.(2) appears as a simple first-order differential equation, yielding $r$ as a function of $t$. However, complexity of solving this differential equation becomes immediately apparent because we will be dealing with the case in which $R(t,r)$ is not factorized into a separable form of $a(t)f(r)$. The boundary condition to be imposed is as follows. Since we measure a signal at $r=0$, the boundary condition is $r(t=t_{received})=0$. More specifically, we treat the solution of Eq.(2) as $r = r(t,t_0)$ which is a function of $t$ with the boundary condition $r(t=t_0, t_0)=0$. In order to define the redshift, we consider two successive wave crests, both of which leave $r$ and reach us ($r=0$) at different times. Suppose that two wave crests were emitted at time $t$ and $t+\Dt$ and received by us at time $t_0$ and $t_0+\Dtz$, respectively. Then, from Eq.(2) and the definition of $r(t,t_0)$, we have \beqarr r & =& \int_t^{t_0} \frac{\di t'}{R(t', r(t',t_0))} \\ \no & =& \int_{t+\Dt}^{t_0+\Dtz} \frac{\di t'}{R(t',r(t',t_0+\Dtz))} {}~. \eeqarr It is to be noted that for each wave crest a proper boundary condition has to be applied, which is explicitly expressed in the form of $r(t,t_{received})$. Since $\Dt$ and $\Dtz$ are extremely small compared with the cosmological time scale, it is sufficient to consider only up to the first order in $\Dt$ or $\Dtz$. Then, we have, from Eq.(3), \beq \frac{ \Dt}{R(t,r(t,t_0))} = \frac{\Dtz}{R(t_0, r(t_0,t_0))} \left[ 1+R(t_0,0)\int_t^{t_0} \di t' \frac{\rd}{\rd t_0} \left( \frac{1}{R(t',r(t',t_0))} \right) \right]~~, \eeq yielding the defining relation of the redshift, $z$, as \beq 1+z \equiv \frac{\Dtz}{\Dt} =\frac{R(t_0,0)}{R(t,r(t,t_0))} \left[ 1+R(t_0,0) \int_t^{t_0}\di t' \frac{\rd}{\rd t_0}\frac{1}{R(t',r(t',t_0))} \right]^{-1}~~, \eeq where the boundary condition, $r(t_0,t_0)=0$ has been used. Here and hereafter ${\rd \over \rd t_0}$ explicitly means $\frac{\rd r}{\rd t_0} \frac{\rd}{\rd r}$. In the case of the Robertson--Walker metric where $R(t,r)\equiv S(t)$, the scale factor of the standard Friedmann cosmology, it is easy to see that Eq.(5) reduces to the well-known relation, $(1+z)=S(t_0)/S(t)$. In the real observation, the most important definition of distance is the {\it luminosity distance\/}. As is well known, if a source at comoving distance $r$ emits light at time $t$ and a detector at $r=0$ receives the light at time $t_0$, the luminosity distance, $d_L$, of a source in the standard Friedmann cosmology is \beq d_L = r S(t_0) (1+z) = r S(t) (1+z)^2~~, \eeq where the second equality is due to the relation $S(t_0)=S(t)(1+z)$ in the standard Friedmann cosmology. The luminosity distance in a spherically symmetric but inhomogeneous Universe was first examined by Bondi \cite{Bondi} in 1947. In order to avoid the non-zero pressure, however, two different scale factors were introduced in the metric, as was originally done by Lemaitre, Tolman and Dingle \cite{Lemaitre}, \beq \di \tau^2 = \di t^2 - X^2(t,r) \di r^2 - Y^2(t,r) \di \Omega^2~~. \eeq One of the special features of the TB model is that the pressure is always zero, as originally designed so as to be applicable in the matter dominated era only. The situation considered in \cite{Bondi} is such that a standard source is at the $center$ and an observer at $(t,r,\theta,\varphi)$. As was shown in \cite{Bondi}, the ratio of the absolute luminosity to the apparent luminosity is simply given by $Y^2(t,r)(1+z)^2$. Two comments are in order here. First, we note that the light source is at the center, implying that the light propagates out spherically with constant surface energy density at any given time, which is, in general, not the case in an inhomogeneous Universe. Secondly, $(t,r,\theta,\varphi)$ is a coordinate of the observer, not of the source. That is, in the standard cosmology notation, $Y(t,r)$ physically corresponds to $rS(t_0)$. Since we consider the situation in which the position of the observer is located at the center, the light from its source off the center does not even propagate outward in a spherically symmetrical manner. Following the picture of the Universe in \cite{Kim} where the Universe is inside a bubble with the underdense center (where the observer is located) and with the highly dense shell, the light would feel attraction toward the shell, implying that the path of the light propagation is, in general, not a straight line. Moreover, its energy is not uniformly distributed over the non-spherical shell at any given time. Nevertheless, since the position of the observer is fixed at the center (i.e., $r=0$) and he/she receives the light that propagates on a straight line, we shall use the following definition of luminosity distance \beq d_L \equiv r(t,t_0)R(t_0, r(t=t_0,t_0))(1+z)~~, \eeq where the $(1+z)$ factor comes from the correction factor, $(1+z)^2$, that appears in the relationship between the absolute luminosity and the apparent luminosity (hence only one factor out of $(1+z)^2$ in the distance). One factor of $(1+z)$ in the luminosity relation is due to the decrease of energy because of the redshift, the other factor coming from the increase of the time interval from $\Dt$ to $\Dtz$, which is also just $(1+z)$ by definition. Of course, the above definition has to be justified by performing the coordinate transformation from the observer to the source in the inhomogeneous Universe. This, however, is beyond the scope of this paper and thus, based on its plausibility, we shall assume its validity in this paper. We caution the reader that the luminosity distance should not be simply written as $r(t,t_0)R(t,r(t,t_0))(1+z)^2$, for $R(t,r(t,t_0))$ is not simply given by $R(t_0,0)$ times a factor, $(1+z)$, as can be seen in Eq.(5). \section{Scale--dependent cosmology} In order to obtain some specific results of cosmological consequences of the proposed inhomogeneous metric, Eq.(1), we shall consider, as an example, the model of \cite{Kim}. In this Section, we shall briefly summarize the model. First, given the metric in Eq.(1), in order to accommodate the $r$-dependence on the Ricci tensors, the Einstein equation was also generalized as \begin{equation} R^{\mu \nu}-\frac{1}{2}g^{\mu \nu}R=-8\pi [G T^{\mu \nu}](t,r)~, \end{equation} where the $(t,r)$ dependence of the combination, $[GT^{\mu \nu}]$, was explicitly noted. When the non-vanishing elements of Ricci tensor calculated from Eq.(1) are substituted into the generalized Einstein equation in Eq.(9), we obtain the following non-vanishing components. \begin{eqnarray} 3 \frac{\dot{R}^2(t,r)}{R^2(t,r)}-2\frac{R''(t,r)}{R^3(t,r)} +\frac{R'^2(t,r)}{R^4(t,r)}-4\frac{R'(t,r)}{rR^3(t,r)} &=&8 \pi [G\rho] \\ 2 \frac{\ddot{R}(t,r)}{R(t,r)}+\frac{\dot{R}^2(t,r)}{R^2(t,r)} -\frac{R'^2(t,r)}{R^4(t,r)}-2\frac{R'(t,r)}{rR^3(t,r)} &=& -8 \pi [Gp_r] \\ 2 \frac{\ddot{R}(t,r)}{R(t,r)}+\frac{\dot{R}^2(t,r)}{R^2(t,r)} -\frac{R''(t,r)}{R^3(t,r)} +\frac{R'^2(t,r)}{R^4(t,r)}-\frac{R'(t,r)}{rR^3(t,r)} &=&- 8 \pi [Gp_{\theta}] \\ 2 \frac{\ddot{R}(t,r)}{R(t,r)}+\frac{\dot{R}^2(t,r)}{R^2(t,r)} -\frac{R''(t,r)}{R^3(t,r)} +\frac{R'^2(t,r)}{R^4(t,r)}-\frac{R'(t,r)}{rR^3(t,r)} &=&- 8 \pi [Gp_{\varphi}] \end{eqnarray} where dots and primes denote, respectively, derivatives with respect to $t$ and $r$. Another non-vanishing Ricci tensor $R_{01}$ yields \begin{equation} R_{01} = 2 \left( \frac{\dot{R}'(t,r)}{R(t,r)} -\frac{\dot{R}(t,r)R'(t,r)}{R^2(t,r)} \right) = - 8\pi [GT_{01}]~~. \end{equation} As was discussed in \cite{Kim}, in order to maintain an inhomogeneous matter distribution, it is essential to keep pressures and $T_{01}$ to be finite so that $R(t,r)$ is kept from being factored out as $a(t)f(r)$, in which case the Robertson--Walker metric is recovered. This feature distinguishes this model from the TB model, in which pressure was set to be zero to begin with. Moreover, to avoid the sheer force, it was assumed in \cite{Kim} that $p_r=p_{\theta}(=p_{\varphi})$. Then a constraint on $R(t,r)$ is uniquely determined from Eqs.(11) and (12) as \beq R(t,r) = \frac{ a(t) }{1-B(t)r^2}~~, \eeq where $a(t)$ and $B(t)$ are positive, arbitrary functions of $t$ alone. The negative sign on the right hand side of Eq.(15) is chosen to avoid a locally closed Universe (see below Eq.(16)). Inserting Eq.(15) into Eq.(10) gives \begin{equation} \left[ \frac{\dot{R}(t,r)}{R(t,r)} \right]^2 = \frac{8\pi }{3}[G\rho](t,r) + \frac{4B(t)} {a^2(t)} ~~. \end{equation} The term, $4B(t)/a^2(t)$, was interpreted as a time-varying vacuum energy density in \cite{Kim}. It should be noted here that the appearance of this term, admittedly very surprising, is a consequence of the metric in Eq.(1). Let us briefly discuss physical implications of Eq.(16). In our neighborhood (i.e., $r \ll 1$), the left-hand side of Eq.(16) is reduced to $(\dot{a}/a)^2$, implying that $a(t)$ represents more or less the scale factor for our local Universe. Moreover, since the standard Friedmann cosmology has been successful in describing our local neighborhood, any modifications of the standard cosmology in this model must be small in the local Universe. The next question is whether our local neighborhood is flat or open. If it is $flat$, $B(t)$ should be treated, as can be seen on the right--hand side of Eq.(16), as being small in accord with the assumption of small modifications on the local neighborhood. If the local neighborhood is $open$, however, $B(t)$ itself is not small. Upon writing $B(t)$ as $[1+b(t)]/4$, Eq.(16) becomes \begin{equation} \left[ \frac{\dot{R}(t,r)}{R(t,r)} \right]^2 = \frac{8\pi }{3}[G\rho](t,r) + \frac{1} {a^2(t)} + \frac{ b(t)}{a^2(t)}~~, \end{equation} implying that in the local neighborhood (i.e., for $r \ll 1$, or equivalently for $(\dot{R}/R)^2 \simeq (\dot{a}/a)^2$), small is $b(t)$, but not $B(t)$. It is interesting to mention here that the metric given by Eq.(1) with the Einstein equation dictates the behavior of the energy density, pressure and momentum density of the Universe. We first note that the constraint on $R(t,r)$, Eq.(15), severely restricts the behavior of $G\rho$, $Gp$ and $GT_{01}$ as functions of $t$ and $r$. Substituting Eq.(15) into Eqs.(10), (11) and (14), we obtain the following explicit expressions : \begin{eqnarray} \frac{8\pi}{3}G\rho &=& \left(\frac{\dot{a}}{a} \right)^2 -\frac{1}{a^2}-\frac{ b}{a^2} + 2\left(\frac{\dot{a}}{a}\right)\left( \frac{\dot{b}r^2/4} {1- \left[ {1+b \over 4} \right]r^2}\right) +\left(\frac{\dot{b}r^2/4} {1- \left[ {1+b \over 4} \right]r^2} \right)^2 \\ 8\pi Gp&=& \frac{1}{a^2}- 2\frac{\ddot{a}}{a} - \left( \frac{\dot{a}}{a}\right)^2 +\frac{b}{a^2} - \left( 6 \frac{\dot{a}}{a}+ 2\frac{\ddot{b}}{\dot{b}} \right) \left( \frac{\dot{b}r^2/4} {1- \left[ {1+b \over 4} \right]r^2}\right) - 5 \left( \frac{\dot{b}r^2/4} {1- \left[ {1+b \over 4} \right]r^2} \right)^2 \\ 8\pi GT^{01} &=& \frac{\dot{b}r}{a^2} ~~. \end{eqnarray} The following comments are in order here. {}From Eqs.(18) and (19), it is easy to see that as $ r$ approaches $\sqrt{\frac{4}{1+b}}$, we have $p/\rho = -\frac{5}{3}$, implying a bizarre equation of state. In the standard inflationary scenario in which a constant vacuum energy is responsible for generating an inflationary period, one has $ p/\rho = -1$. Therefore, in the model under consideration, an unusual scalar sector with time-dependent vacuum energy, which is as yet to be understood, may be responsible for the inflation which is much more rapid than the usual inflation. By using the arbitrariness of $b(t)$, however, this singularity with the unusual $p/\rho$ ratio can be pushed far away from the particle horizon, so that the local value of $p/\rho$ at the present matter dominated era can be made to be positively small, as generally expected. For this reason, in spite of this bizarre behavior, we shall proceed to use this model as an example for our following discussions. \section{Perturbative Approach} The linear relationship between the redshift and luminosity distance with a constant coefficient (the Hubble constant) for nearby objects (for $z \ll 1$) has been well established by various observations. In this Section, we investigate whether or not the redshift--luminosity distance relation for small $z$ in an inhomogeneous cosmological model discussed in the previous Section still remains the same as in the standard cosmology. \subsection{Locally flat Universe} We start with the light propagation equation \beq \frac{\di r}{\di t}=-\frac{1}{a(t)}+\frac{B(t)}{a(t)}\,r^2, \eeq which is determined by two arbitrary functions of $t$, $a(t)$ and $B(t)$. But unfortunately Eq.(21) cannot be solved analytically because it is non-linear in $r$. For a locally $flat$ Universe with small perturbations to the standard cosmology, $B(t)$ should be treated as being very small. Therefore, we have, from Eq.(21), \beq r(t,t_0)= \int_t^{t_0}\frac{\di t'}{a(t')} -\int_t^{t_0} \di t' \frac{B(t')}{a(t')} \left[ \int_{t'}^{t_0}\frac{ \di t''}{a(t'')} \right]^2 +{\cal O\/}(B^2)~~, \eeq yielding \beq \frac{\rd}{\rd t_0}\left[ \frac{1}{R(t,r(t,r_0))} \right]= -\frac{2 B(t)}{a(t)a(t_0)} \int_t^{t_0}\frac{\di t'}{a(t')} + {\cal O\/}(B^2)~~, \eeq where we have used Eq.(15). Now, the redshift is given by \beqarr 1+z &\equiv& \frac{\Dtz}{\Dt} \\ \no &=& \frac{a(t_0)}{a(t)} \left[ 1-B(t)\left( \int_t^{t_0}\frac{\di t'}{a(t')} \right)^2 +2\int_t^{t_0}\di t' \frac{B(t')}{a(t')} \int_{t'}^{t_0}\frac{ \di t''}{a(t'')} + {\cal O\/}(B^2) \right]~~. \eeqarr In a special case where $B(t)$ is a constant, which corresponds to the standard Friedmann cosmology, the redshift simply reduces to the standard relation \beq 1+z = \frac{a(t_0)}{a(t)}~~, \eeq where we have used the relation \beq 2 \int_t^{t_0}\frac{\di t'}{a(t')} \int_{t'}^{t_0}\frac{ \di t''}{a(t'')} = \left[ \int_t^{t_0}\frac{\di t'}{a(t')} \right]^2 ~~ . \eeq That is, the well-known redshift expression in the standard cosmology is reproduced, as expected. Therefore, this result strongly suggests that $a(t)$ plays a role of, more or less, the scale factor of the standard Friedmann cosmology. In the case of a locally flat Universe, therefore, the behavior of $a(t)$ cannot be much different from that of the standard Friedmann cosmology with $k=0$, which is proportional to $t^{2 \over 3}$ in the matter dominated era. For mathematical simplicity and illustrative purposes, we assume that $a(t)$ behaves the same as that in the standard cosmology and $B(t)$ can be expressed by a simple power law. That is, we assume that $a(t)$ and $B(t)$ are, in the matter dominated era, of the form \beq a(t)=\alpha t_0 \, \left( \frac{t}{t_0} \right) ^{ {2 \over 3} } {}~,~~ B(t)=\beta \left( \frac{t}{t_0} \right)^n {}~( n \ge 0)~~, \eeq where $\alpha$ and $\beta$ are dimensionless parameters to be determined and $n$ is set to be non-negative because of the observed increase of the matter density as a function of $r$.\cite{Kim} Substituting Eq.(27) into Eq.(22) yields \beqarr r(t,t_0) &=& \frac{3}{\alpha} \left[ 1- \left( \frac{t}{t_0} \right)^ {1 \over 3} \right] -\beta\frac{9}{\alpha^3} \left[ \frac{2}{9 ( n+ {1 \over 3})(n+{2 \over 3})(n+1)}\right. \\ \no & &\left.- \frac{1}{(n+{1 \over 3})} \left( \frac{t}{t_0} \right)^{ n+{1\over 3}} +\frac{2} { (n+ {2 \over 3}) } \left( \frac{t}{t_0} \right)^{n+{2\over 3}} -\frac{1}{(n+1)} \left( \frac{t}{t_0} \right)^{n+1} \right] +{\cal O\/}(\beta^2)~~. \eeqarr In practice, what is measured is the redshift. Therefore, we must express $( t/t_0)$ in terms of the red shift. By substituting Eqs.(27) into Eq.(24), we have \beqarr 1+z &=& \left( \frac{t}{t_0} \right)^{- {2 \over 3}}\left[ 1-\beta \frac{9}{\alpha^2}\left( \frac{t}{t_0} \right)^n \left\{1- \left( \frac{t}{t_0} \right)^{1\over 3} \right\}^2 \right. \\ \no & & \left. + \beta \frac{2}{ \alpha^2(n+ {1 \over 3})(n+ {2 \over 3}) } \left\{ 1+3 (n+{1 \over 3}) \left( \frac{t}{t_0} \right)^{n+{2 \over 3}}-3 (n+ {2 \over 3}) \left( \frac{t}{t_0} \right)^{n+ {1 \over 3}}\right\} \right] + {\cal O\/}(\beta^2)~~, \eeqarr yielding \beqarr \left( \frac{t}{t_0} \right) &\equiv& T_{(0)}+\beta T_{(1)}+{\cal O\/}(\beta^2)~~, \eeqarr where, for notational simplicity, we define $T_{(0)}(z)$ and $T_{(1)}(z)$ as \beqarr T_{(0)}&=& (1+z)^{- {3 \over 2}} \\ \no T_{(1)} &=&{3 \over 2\alpha^2} \left[ \frac{ 2 }{(n+ {1 \over 3})(n+ {2 \over 3})} (1+z)^{- {3 \over 2}}+ \frac{18n}{(n+ {1 \over 3})}(1+z)^{- {3 \over 2}(n+{4\over 3})} \right. \\ \no & & \left. -\frac{9n}{(n+ {2 \over 3})}(1+z)^{- {3 \over 2}(n+{5 \over 3})} -9(1+z)^{-{3 \over 2}(n+1)}\right]~~. \eeqarr Inserting Eqs.(30) and (31) into the right--hand side of Eq.(28) gives \beqarr r(t,t_0) & = & r_{(0)}+\beta r_{(1)} +{\cal O\/}(\beta^2)~~, \eeqarr where $r_{(0)}$ and $r_{(1)}$ are defined as \beqarr r_{(0)} &\equiv& {3 \over \alpha} [1- T_{(0)}^{1 \over 3}] \\ \no r_{(1)} &\equiv& -{1 \over \alpha^3} \left[ T_{(0)}^{-{2 \over 3}} T_{(1)} \alpha^2 + \frac{2}{ (n+ {1\over3})(n+{2 \over 3})(n+1)} \right. \\ \no & &\left. -\frac{9}{(n+ {1 \over 3})} T_{(0)}^{ n+{1 \over3}} +\frac{18}{(n+{2 \over 3})} T_{(0)}^{n+{2 \over 3}} -\frac{9}{(n+1)} T_{(0)}^{n+1} \right]. ~~. \eeqarr It is easy to see that the $r_{(0)}$ term reproduces the result in the standard Friedmann cosmology whereas $r_{(1)}$ represents a correction term. It is interesting to note that for small $z$, $r_{(1)}$ is zero up to the second order in $z$, as can easily be seen by substituting Eq.(31) into Eq.(33). That is, there is no modification of small $r(t)$ (i.e., for $z\ll 1$) due to small $B$, up to the second order. Since the luminosity distance is $r(t,t_0)R(t_0,0)(1+z)$, the redshift--luminosity distance relation for small $z$ remains intact, at least, up to the first order in $B(t)$. This is a consequence of the plausible assumption that $a(t)$ has very similar behavior of the scale factor of the standard Friedmann cosmology in the matter dominated era. \subsection{Locally Open Universe.} Before we proceed, we make a brief comment on the status of the density parameter $\Omega$. One of the most challenging tasks in the observational astrophysics is the measurement of the mass density of the Universe, which is supposed to be a constant at any given time in the standard Friedmann cosmology. Various observations, however, indicate that the mass density indeed appears to increase as we probe farther out \cite{Schramm}\cite{Kolb}. {}From direct observations, the fraction of critical density associated with luminous galaxies is $\Omega_{LUM}\leq 0.01$. When extending the observation to distances beyond the luminous part of galaxies, we found that there exist galactic halos which have a mass corresponding to $\Omega_{HALO} \simeq 0.1$. On a larger scale such as the Virgo cluster, modeling the local distortion of the Hubble flow around the cluster yields $\Omega_{CLUSTER}=$0.1 to 0.2. Recently, using the redshift measurements for the catalogue of galaxies by the Infrared Astronomy Satellite (IRAS)\cite{IRAS}, it became apparent that galaxies out to about 100 Mpc flow towards the Great Attractor with high peculiar velocity. It was concluded that the observed dynamics on this scale requires $\Omega_{IRAS} \sim 1 \pm 0.6$. Based on the above observations, we have a picture of the Universe in which our local neighborhood is underdense and the mass density increases with scale. Therefore, we present, in this Subsection, the redshift--luminosity distance relation for the locally $open$ Universe. As was discussed before, $a(t)$ is more or less the scale-factor of the locally open Universe in the standard Friedmann cosmology. Therefore, it is more transparent to rewrite $B(t)$ as $[1+b(t)]/4$. Hereafter, a small perturbation to the locally $open$ Universe is represented by $b(t)$ rather than by $B(t)$. It is convenient to introduce a new coordinate $\Phi$ as defined by $r(t,t_0) = 2 \tanh \Phi(t,t_0)$. The light propagation equation is then reduced to \beq \frac{\di \Phi}{\di t}= -\frac{1}{2a(t)} +\frac{b(t)}{2 a(t)} \sinh^2 \Phi~~. \eeq We consider two successive wave crests that leave a comoving coordinate $\Phi$ at time $t$ and $t+\Dt$ and arrive at $\Phi=0$ at times $t_0$ and $t_0+\Dtz$, respectively, which yields the following equality : \beqarr \Phi &=&\int_{t}^{t_0}\di t' \left[ \frac{1}{2a(t')}-\frac{b(t')}{2a(t')} \sinh^2 \Phi (t',t_0) \right] \\ \no &=& \int_{t+\Dt}^{t_0+\Dtz}\di t' \left[ \frac{1}{2a(t')}-\frac{b(t')}{2a(t')} \sinh^2\Phi(t',t_0+\Dtz) \right]~~, \eeqarr from which the redshift relation is given by \beq 1+z \equiv \frac{\Dtz}{\Dt} = \frac{a(t_0)}{a(t)} \left[ \frac {1-b(t)\sinh ^2 \Phi(t,t_0)} {1-a(t_0) \int_{t}^{t_0}\di t' \frac{b(t')}{a(t')} \frac{ \rd \Phi(t',t_0)}{\rd t_0} \sinh 2\Phi (t',t_0) } \right]. \eeq For an open Friedmann Universe where $b(t)=0$, $(1+z)$ simply becomes $a(t_0)/a(t)$, which is the standard result. Since the light propagation equation is non-linear, we will again use the perturbation method by treating $b(t)$ as being small. Then, the redshift in this picture becomes \beq 1+z \simeq \frac{a(t_0)}{a(t)}\left[ 1-b(t)\sinh^2 \int_{t}^{t_0}\frac{\di t'}{2a(t')} + \int_{t}^{t_0}\di t' \frac{b(t')}{2a(t')}\sinh \int_{t'}^{t_0} \frac{\di t''}{a(t'')} +{\cal O\/}(b^2)\right]~~, \eeq where we have used the relation, $\rd \Phi(t,t_0)/\rd t_0 = 1/2 a(t_0)$. Again to proceed further we need specific functional forms of $a(t)$ and $b(t)$. Since $a(t)$ cannot be too different from the scale factor in the Friedmann cosmology with $k=-1$, we assume that, in the matter-dominated era, $a(t)$ satisfies the following differential equation as in the standard Friedmann cosmology with $k=-1$ : \beq 2\, \frac {\ddot{a}(t)}{a(t)} + \left( \frac{\dot{a}(t)}{a(t)}\right)^2 -\frac{1}{a^2(t)} \simeq 0 ~~. \eeq Here, only one of the two initial conditions can be fixed as $a(t=0)=0$. The solution of Eq.(38) may be parameterized by an angle, $\Psi$, as \beqarr a(\Psi) &=& \alpha t_0 [\cosh \Psi -1] \\ \no t(\Psi) &=& \alpha t_0 [\sinh \Psi -\Psi] {}~, \eeqarr where $t_0$ is the age of the Universe and $\alpha$ is a dimensionless parameter to be determined. In the following, $\Psi_0$ is defined as $t(\Psi_0)=t_0$, which would correspond to $(1-q_0)/q_0$ in the Friedmann cosmology with $k=-1$, where $q_0$ is the deceleration parameter. For illustrative purposes, we consider, in this Section, a simple case where the arbitrary function $b(t)$ behaves as \beq b(\Psi) = \beta \Psi~~, \eeq where $\beta$ is a dimensionless parameter to be treated as a perturbation. Then, from Eqs.(35) and (37), we have \beq \frac{a(t)}{a(t_0)} = \frac{1}{1+z} \left[ 1+\frac{\beta}{2} \{ \sinh(\Psi_0-\Psi)-(\Psi_0-\Psi)\} \right] + {\cal O\/}(\beta^2) \eeq and \beqarr \Phi(\Psi,\Psi_0) &=& \frac{1}{2} (\Psi_0-\Psi) \\ \no & & -\frac{\beta}{4} \left[ \cosh (\Psi_0-\Psi) +\Psi\sinh (\Psi_0-\Psi)-\frac{\Psi_0^2-\Psi^2}{2}-1\right] + {\cal O\/}(\beta^2)~~. \eeqarr Now, we are ready to obtain the redshift--luminosity relation based on $d_L=r R(t_0,0)(1+z)$. Recalling $r(t,t_0)=2\tanh\Phi(t,t_0)$, $(\Psi_0-\Psi)$ should be expressed in terms of the redshift. {}From the parameterized solution Eq.(39), $\Psi$ is \beq \Psi = \cosh^{-1} \left[ 1+( \cosh \Psi_0-1 )\frac{a(t)}{a(t_0)} \right]~~. \eeq Substituting Eq.(41) into Eq.(43), we find \beqarr \Psi_0-\Psi &=& \Psi_0-\cosh^{-1} \left[1+\frac{ \cosh \Psi_0-1}{1+z} \right] \\ \no & & -\frac{\beta}{2} \sqrt{\frac{ \cosh \Psi_0-1} {1+\cosh\Psi_0 + 2 z} } \left\{ \sinh (\Psi_0-\Psi)- (\Psi_0-\Psi) \right\} + {\cal O\/}(\beta^2)~~. \eeqarr For simplicity, we define the zeroth order term of $\Psi$, $\Psi^{(0)}$, as \beq \Psi^{(0)} = \cosh ^{-1} \left[ 1+\frac{ \cosh{\Psi_0}-1}{1+z} \right]~~. \eeq Then, from Eqs.(42), (44) and (45), we have \beqarr \Phi(\Psi,\Psi_0) &=& \frac{1}{2} (\Psi_0 - \Psi^{(0)}) \\ \no & & + \frac{\beta}{4} \sqrt{\frac{ \cosh \Psi_0-1} {1+\cosh\Psi_0 + 2 z} } \left[ (\Psi_0 - \Psi^{(0)}) - \sinh (\Psi_0 - \Psi^{(0)}) \right] \\ \no & & - \frac{\beta}{4} \left[ \cosh (\Psi_0 - \Psi^{(0)}) +\Psi^{(0)}\sinh (\Psi_0 - \Psi^{(0)}) -\frac{ \Psi_0^2-(\Psi^{(0)})^2}{2}-1 \right] \\ \no & & +{\cal O\/} (\beta^2) \\ \no &\equiv& \Phi^{(0)}(z)+\frac{\beta}{4} \Phi^{(1)} + {\cal O\/} (\beta^2)~~. \eeqarr In our neighborhood ($z\ll 1$), $\Phi^{(0)}(z)$ becomes \beq \Phi^{(0)} = \frac{1}{2} \sqrt{ \frac{\cosh \Psi_0-1}{\cosh \Psi_0 +1}} \left[ z - \frac{ \cosh \Psi_0+2}{2(\cosh\Psi_0+1)} \,z^2 \right] + {\cal O\/}(z^3)~~. \eeq Using Eqs.(44), (45) and (46), it can easily be seen that $\Phi^{(1)}$ is zero up to the second order of $z$. Finally, $r(t,t_0)$ is expressed in terms of the redshift by \beqarr r(t,t_0) &=& 2 \tanh \Phi(t,t_0) \\ \no & = & 2 \tanh \Phi^{(0)}(z) +\frac{\beta}{2}\Phi^{(1)}(z) (1-\tanh^2 \Phi^{(0)}(z)) +{\cal O\/}(\beta^2)~~ \\ \no &=& \sqrt{ \frac{ \cosh \Psi_0 -1}{\cosh \Psi_0+1} } \left[ z-\frac{ \cosh \Psi_0+2}{2(\cosh \Psi_0+1)} \, z^2 \right] + {\cal O\/}(z^3, \beta^2)~~. \eeqarr If $\cosh \Psi_0=(1-q_0)/q_0$ as in the Friedmann cosmology, the standard result, $r = \sqrt{1-2q_0}[z-(1+q_0)z^2/2]+{\cal O\/}(z^3)$, is recovered, implying that the luminosity distance, $r R(t_0)(1+z)$, has the same dependence on $z$ for the small redshift as in the standard model. To summarize, the standard results of the redshift--luminosity distance relation in the Friedmann cosmology for $z \ll 1$ are preserved in this model, regardless of whether the local Universe is assumed to be open or flat. \section{Numerical Results} As was demonstrated in the previous section, the most fundamental cosmological relation predicted by this model, i.e., the redshift--luminosity relation, is practically the same as in the standard cosmology for small $z$. This is due to our plausible assumption that $a(t)$ behaves more or less the same as the scale factor in the standard cosmology. We now proceed to explore the relationship for large $z$. Unfortunately, it is impossible to do so analytically, due partly to our inability to analytically solve the light propagation equation. In this Section, therefore, we resort to numerical method and present some numerical answers in the form of figures depicting possible modifications of the redshift--luminosity distance relation for large $z$. Since our local Universe appears to be open as discussed before, we only consider the second case of the previous section. Thus, $a(t)$ is parameterized by an angle $\Psi$ as given in Eq.(39). Since $b(t)$ is completely arbitrary except for being small, we consider the following three representative cases : (A) $b(\Psi)=\beta \Psi$, (B) $b(\Psi)=\beta \Psi^4$ and (C) $b(\Psi)=\beta [\cosh \Psi - 1]^2$, where $\beta$ is a dimensionless parameter to be determined. Note that the case (C) corresponds to the picture in which there is a $constant$ vacuum energy density in the Universe, as can be seen from Eq.(17). Recalling that the Hubble expansion rate of the local Universe at the present epoch, $\overline{H}_0\equiv \dot{R}/R|_{t=t_0,r\sim0}$, is $[ \dot{a}(t)/a(t)]_{t=t_0}$, the density parameter in our neighborhood is, from Eq.(18), \beq \overline{\Omega}_0 \equiv \frac{8 \pi G \rho(t=t_0,r\simeq 0)} {3 \overline{H}_0^2} = 1- \dot{a}_0^2 - b(t_0)\dot{a}_0^2~~, \eeq where the bar denotes the $local$ value and the subscript zero represents the $present$ value. With the definitions of $ \overline{\Omega}_{0,a} \equiv 1-\dot{a}_0^2$ and $\overline{\Omega}_{0,b} \equiv b_0 \dot{a}_0^2 $, the density parameter and the pressure of the local Universe can be expressed by \beqarr \overline{\Omega}_{0} &=& \overline{\Omega}_{0,a} -\overline{\Omega}_{0,b}~~, \\ 8 \pi\overline{ Gp_0} &=&\overline{\Omega}_{0,b} \overline{H}_{0}^2~~, \eeqarr where Eq.(19) is used. {}From the observation of small peculiar velocities of nearby galaxies, we assume that the pressure of the local Universe is relatively small compared with the energy density. That is, $\overline{\Omega}_{0,b} $ is very small. In the numerical calculations to be presented in this Section, we assume $\overline{\Omega}_{0} =0.1$ and, for the sake of definiteness, choose values $\overline{\Omega}_{0,a} =0.1001$ and $\overline{\Omega}_{0,b} =0.0001$, specifying the numerical values of $\alpha$ and $\beta$. The light propagation equation can then be numerically integrated with the boundary condition $r(t=t_0, t_0)=0$. To this end, we divide the numerical solution of $r(t,t_0)$ into $N$ intervals and fit each interval using \beq r_i(t,t_0)= \delta_i \left[ t_0^{\gamma_i} -t^{\gamma_i} \right]~~\mbox{ for\/}~~ \frac{i-1}{N} < \frac{t}{t_0} < \frac{i}{N}~~ (i=1,2,3...)~, \eeq which then yields the numerical values of $\delta_i$ and $\gamma_i$ for each interval. We have used $ N = 200$ in our numerical calculations. Using the definition of the comoving distance, i.e., $r_i(t,t_0)=r_i(t+\Dt,t_0+\Dtz)$, we have the redshift, for each interval, as \beq (1+z) \equiv \left( \frac{\Dtz}{\Dt} \right)_i = \left( \frac{t}{t_0} \right)^{\gamma_i-1} {}~. \eeq Then, the luminosity distance, $d_L=r R(t_0)(1+z)$, for each interval, can easily be calculated from Eqs.(52) and (53). In Fig.1, the results of the three cases discussed above are presented by solid lines, while the standard results with $\Omega=0.1$ and $\Omega=1.0$ by the dashed lines and dotted lines, respectively. First, it is to be noted that on small scale, the linear {\em Hubble diagrams\/} are preserved for all of the three cases, as was shown by perturbative calculations. In the cases of (A) and (C), the redshift--luminosity distance relations in this model are almost indistinguishable and furthermore correspond to the standard results with the density parameter in the range $\Omega \sim$ $0.8-0.9$, even though the local mass density in the model is set to be $\overline{\Omega}_0=0.1$. In the case (B), however, it appears to be the standard result with $\Omega > 1$ although the difference is not very significant. Nevertheless, this qualitative feature was totally unexpected. In this model, therefore, a precise measurement of the redshift--luminosity distance relation alone cannot provide information on $q_0$, which is related to $\Omega_0$ as $2q_0=\Omega_0$, contrary to the case of the standard cosmology. Now it would be appropriate to discuss the meaning of the $observed$ increase of $\Omega_0$ as we look farther out. Every light signal we are receiving right now contains information about the past in time. That is, what we measure are, for example, the redshift and $G\rho(t)$, not $G\rho(t_0)$. Thus, we deduce physical quantities at the present time $t_0$, using the standard cosmological evolution equations. Recalling that $G\rho(t)$ in the matter dominated era in the standard Friedmann cosmology is proportional to $1/S^3(t)$, and $S(t)$ is just $S(t_0)/(1+z)$, $G\rho(t_0)$ is $obtained$ from $G\rho(t)$ by multiplying $1/(1+z)^3$. Dividing $G\rho(t_0)$ with the constant critical density in the standard cosmology, $G\rho^{(s)}_c$, we can deduce the density parameter at present time, which we shall call $\Omega^{obs}_0$, as \beq \Omega^{obs}_0 \equiv \frac{G\rho^{obs}(t_0)}{G\rho^{(s)}_c} \equiv \frac{G\rho(t,r(t))}{G\rho^{(s)}_c (1+z)^3}~~. \eeq Using Eqs.(52) and (53) for each interval in Eq.(18), we have calculated $\Omega_0^{obs}$ versus the redshift. The results are shown in Fig.2. As before, we have taken the local value $\Omega_{0}^{obs} =0.1$. In all three cases, the calculated values of $\Omega^{obs}_0$ are increasing functions of $z$. ( A naive value in the standard cosmology is supposed to be a constant.) We can see that the cases of (A) and (C) are practically identical, whereas the case (B) shows a faster increase of $\Omega_0^{obs}$. As can be seen in Fig.2, however, even in the case (B) the increase is too slow to explain the IRAS data\cite{IRAS} where $\Omega^{obs}_0$ is compatible to unity (but with large errors) at the distance of about several 100 Mpc($z \sim 0.0166$). To fit the increase of $\Omega^{obs}_0$ up to unity at $\sim 100$ Mpc requires drastic ( perhaps unrealistic ) changes in the form of and parameters in $a(t)$ and $b(t)$, which then would modify our predictions on the Hubble law. \section{summary and conclusions} We have studied how cosmological observables are modified in an isotropic but inhomogeneous Universe compared with those of the standard model. In particular, the luminosity distance and the density parameter as functions of the redshift have been examined in the generalized Robertson--Walker spacetime with only one $(t,r)$-dependent scale factor and they were compared with the standard results. When $R(t,r)$ is not factorized into the form of $a(t)f(r)$, the simple redshift--scale factor relation such as $(1+z)=a(t_0)/a(t)$ remains no longer valid. First by solving light propagation equation, Eq.(2), for radially propagating light with the boundary condition $r(t=t_{received})=0$ and then considering two wave crests emitted at time $t$ and $t+\Dt$ which are received at $t_0$ and $t_0+\Dtz$, respectively, we have obtained the general redshift--scale factor relation given by Eq.(5). The result is valid in an inhomogeneous Universe and is shown to be reduced to the simple form $(1+z)=a(t_0)/a(t)$ in the case of the homogeneous spacetime, i.e., the standard Robertson--Walker spacetime. Our general relations agree with the results obtained in \cite{Bondi} and \cite{Moffat}. We have applied the general redshift--scale factor relation to the cosmological model in \cite{Kim} where the Universe is pictured as being inside a highly dense and rapidly expanding shell with the underdense center. First, for the nearby objects ($z\ll 1$), the luminosity distances as functions of the redshift are obtained analytically, using a perturbative method for two cases where the underdense center is either flat or open according to the definition of the standard Friedmann cosmology. One of the most interesting features in \cite{Kim} is that the scale factor $R(t,r)(=a(t)/(1-B(t)r^2) )$ is specified by two arbitrary functions, $a(t)$ and $B(t)$ (or $b(t)$), and $a(t)$ is very similar to the scale factor of the standard Friedmann cosmology and $B(t)$ (or $b(t)$) is the perturbation to the locally flat (or open) Universe. Under the assumption that $a(t)$ behaves the same as that in the standard cosmology, it is shown analytically that the standard redshift--luminosity distance relations in the Friedmann cosmology for $z \ll 1$ remains intact for both cases. Specifically, since the corrections of order $\cal{O \/}(\beta)$ of these expressions can be expanded as a power series of $\Sigma_{i=3}^{\infty} c_i z^i$ with some coefficients $c_i$ (that is, zero up to the second order of $z$), it has been shown that for nearby objects, in spite of its different metric, the cosmological model of \cite{Kim} is not much different from the standard cosmology as far as the Hubble law is concerned. It is also interesting to note that in spite of the special location of the observer, i.e., the return of the pre--Copernican notion in the model \cite{Kim}, the results are almost the same as those of the standard model. As for large $z$, the redshift--luminosity distance relations given in Eqs. (32) and (48) are different from those of the standard cosmology and moreover it is expected that the corrections would be larger when ${\cal O\/}(\beta^2)$ terms are included. In this case, as we mentioned repeatedly, we cannot use the perturbative method and thus we have solved them numerically and obtained the results as shown in Fig.1. and Fig.2. Figure 1 shows the redshift--luminosity distance relation in the cosmological model of \cite{Kim}. Comparing them with the standard cosmology with $k=0$ (dotted curve) and $k=-1$ (dashed curve), we can easily see that for small $z$, the redshift--luminosity distance relation of model \cite{Kim} denoted by the solid lines is almost the same as the standard one, as was also shown in the explicit perturbative calculation. But for $1<z<3$, the Hubble law of the model is very similar to that of the standard cosmology with $k=0$, not with that with $k=-1$, in spite of the fact that the mass density of the local Universe is set to be $\overline{\Omega}_0=0.1$. It is to be noted that although there is no substantial deviation from the standard model in our redshift--luminosity distance relation in Fig.1, the deviation in the case of the TB model is more prominent and is different from ours, both qualitatively and quantitatively as was shown in \cite{Moffat}. Figure 2 shows that the calculated density parameter at the present time, $\Omega^{obs}_0$, is an increasing function of the redshift. Although, admittedly, our perturbative calculations may not be rigorous in the sense that the functions $a(t)$ and $B(t)$ that appear in the scale factor $R(t,r)$ could not be determined by the equation of state, and furthermore our numerical results can only explain the IRAS data qualitatively, we feel that the qualitative nature of our results are robust because we have considered the general case with $B(t)=\beta (t/t_0)^n ~(n >0)$, $b(t)=\beta \Psi(t)$, $b(t)=\beta \Psi^4(t)$ and $b(t)=\beta [ \cosh \Psi -1]^2$ with a variety of numerical values of the parameters involved, all of which have led to similar results. In summary, although the scale-dependent cosmology for the inhomogeneous Universe as modeled in \cite{Kim} implies the explicit running of $H_0$, $\Omega_0$ and $t_0$ as functions of $r$ because of the non-Robertson-Walker metric, as far as the observables such as redshift--luminosity distance relations are concerned, the results are hardly different from those of the standard model in our neighborhood, i.e. for small $z$. Even for large $z$, the difference between the model considered and the standard model with $k=0$ still remains small but the model can be tested when the data from galaxy redshift survey at long distance become available and are compared with the predictions of the model on the matter distribution and on the age of the Universe.\cite{Kim} \acknowledgments The authors would like to thank G. Feldman, M. Im and Y.C. Pei for helpful discussions. One of the authors (THL) wishes to thank the Department of Physics and Astronomy, the Johns Hopkins University for the hospitality extended to him during the completion of this work. This work was supported in part by the Basic Science Research Institute Program, Ministry of Education, Korea, Project No. BSRI-94-2418(THL) and by the National Science Foundation.
\subsection{\bf Introduction} \hspace*{1em} It has been known that the Flavor Changing process in down quark sector is sensitive to the mass difference among up-type quark in the standard model (SM). GIM mechanism tells us that the FCNC in one-loop level of SM vanishes when up type quarks are degenerate. In the real world , there is large mass gap among up quark sector ($M_u \ll M_c \ll M_t$). Then the Flavor Changing Process in down quark sector like $b \rightarrow s \gamma$ and $b \rightarrow s l^+ l^-$ are enhanced. If we extend the fermion sector beyond the standard model, there are several different possibilities. If we just add the chiral fermions as fourth generation in sequential way, GIM mechanism still works and the new contribution comes from up-type quark ($t^{\prime}$) in the fourth generation. Therefore we can get constraints for the mass of $t^\prime$ and its Kobayashi Maskawa mixing to light down-type quarks ($V_{CKM}^{t^\prime i}$, $i = d,s,b$) However if we extend the fermion sector in non-sequential way, the different aspect arises (refs. \cite{branco}, \cite{nir}, \cite{morozumi}, \cite{gautam} and \cite{handoko}). The tree level FCNC arises both in $Z$ and neutral Higgs sector. Furthermore, the size of the FCNC depends on the structure of the down quark mass matrix rather than up type quark masses. Therefore we may study the structure of down type quark mass matrix by studying the flavor changing process in down quark sector. \subsection{\bf Non-sequential extention of fermion sector} \hspace*{1em} We study the Standard Model (SM) with extended quark sector. In addition to the three standard generations of quarks, $N_d - 3$ down-type and $N_u - 3$ up-type vector like singlet quarks are introduced (refs. \cite{branco}, \cite{nir} and \cite{morozumi}). \begin{equation} {I_W} = {1\over 2}\: : \: \left( \begin{array}{c} u \\ d \end{array} \right)_L \left( \begin{array}{c} c \\ s \end{array} \right)_L \left( \begin{array}{c} t \\ b \end{array} \right)_L , I_W = 0 \: : \: \left( \begin{array}{ccccc} u_R & c_R & t_R & t^{\prime}_{L+R} & \cdots \\ d_R & s_R & b_R & b^{\prime}_{L+R} & \cdots \end{array}\right) \: . \end{equation} In this model, there is tree level FCNC in $Z$ and Higgs sector. In order to illustrate this point and explain the relation between the size of FCNC and down-type quark mass matrix, let us introduce a toy model. This is the so called "top-prime less" model \begin{equation} {I_W} = {1\over2} \: : \: \left( \begin{array}{c} {t}^0 \\ {b}^0 \end{array} \right)_L \: , \: {I_W} = 0 \: : \: \begin{array}{c} {t_R}^0 \\ {b_R}^0 \end{array} {b_{L+R}}^{0\prime} \: . \end{equation} The most general mass matrix for down quark sector $M_d$ is given by, \begin{equation} L_{mass} = -(\bar{{b_L}^0} \: \bar{{{b_L}^0}^{\prime}}) \left[ \begin{array}{cc} m & 0 \\ J & m_4 \end{array} \right] \left( \begin{array}{c} {b_R}^0 \\ {{b_R}^0}^{\prime} \end{array}\right) \: . \end{equation} where $J$ is complex and $m$ and $m_4$ are real numbers. $m$ comes from the vacuum expectation value of Higgs doublet. $J$ leads to the mixing between left handed singlet quarks and right handed ordinary quarks. In the limit of $J = 0$, the vector like quark decouples from the ordinary quark. $ M_d {M_d}^\dagger$ can be diagonalized by the following unitary transformation : \begin{equation} \left( \begin{array}{c} {b_L}^0 \\ {b_L}^{\prime 0} \end{array} \right) = \left( \begin{array}{cc} \cos \theta & - \sin \theta e^{-i \phi} \\ \sin \theta e^{i \phi} & \cos \theta \end{array} \right) \left(\begin{array}{c} b_L \\ b_L^{\prime} \end{array} \right) \: , \end{equation} where $b^0$ and $b^{0\prime}$ indicate the weak basis. When $m$ is much smaller than $M = \sqrt{|J|^2 + {m_4}^2}$, the elements of the unitary matrix are given by the following formulae approximately : \begin{eqnarray} \cos \theta & \simeq & 1 \: , \\ \sin \theta e^{-i \phi} & \simeq & {m J \over M^2} \: . \end{eqnarray} Correspondingly, the mass eigenvalues for heavy quark $(M_H)$ and light quark $(M_L)$ are given by : \begin{eqnarray} M_L & = & m {m_4 \over M} \\ M_H & = & M \: . \end{eqnarray} There are two interesting limits : \begin{enumerate} \item $J \ll m_4$ \item $J \gg m_4$ \end{enumerate} In the case of (1), the mixing between ordinary quark and vector like quark is suppressed. The physical masses are given by the diagonal elements of the mass matrix: \begin{equation} M_L = m \: , \: M_H = m_4 \: . \end{equation} In the case of (2), the diagonal elements no longer reflect the physical masses : \begin{equation} M_L = m {m_4 \over |J|} \: , \: M_H = |J| \: . \end{equation} Note that the light quark mass vanishes in the limit $m_4 = 0$. On the other hand, the neutral current is written in terms of physical basis, \begin{eqnarray} L_Z & = & -{g \over 2 \cos \theta_W} \bar{{b^0_L}} \gamma_{\mu} {b^0}_L Z^{\mu} \nonumber \\ & = & -{g \over 2 \cos \theta_W} \left[ \cos^2\theta \bar{b_L} \gamma_{\mu} b_L -\cos \theta \sin \theta e^{-i \phi} \bar{b_L} \gamma_{\mu} b_L^{\prime} + h.c. \right]Z^{\mu} \: . \end{eqnarray} Thus the flavor diagonal coupling for neutral current ($Z^{bb}$) and the FCNC coupling ($Z^{bb^\prime}$) are given by : \begin{eqnarray} Z^{bb} & = &\cos^2\theta = 1 - \sin^2\theta \simeq 1 - O\left( \left(m J \over M^2 \right)^2 \right) \: , \\ Z^{bb^\prime} & = & -\cos \theta \sin \theta e^{-i \phi} \simeq -{m J\over M^2} \: . \end{eqnarray} Depending on the two cases mentioned above, the enhancement and suppression of the FCNC occur respectively. \begin{enumerate} \item $J \ll m_4$ (suppression) \begin{eqnarray} M_L & = & m \: , \\ M_H & = & m_4 \: , \\ \left| Z^{bb^{\prime}}\right| & = & {m |J| \over {m_4}^2} \ll {M_L \over M_H} \: . \end{eqnarray} \item $J \gg m_4$ (enhancement) \begin{eqnarray} M_L & = & m {m_4 \over |J|} \: , \\ M_H & = & \left| J \right| \: , \\ \left| Z^{bb^{\prime}} \right| & = & {m \over |J|} \gg {M_L \over M_H} \: . \end{eqnarray} \end{enumerate} For instance, if the physical mass of the vector like quark mass is $M_H = 500$(GeV), the following mass matrices realize the same mass eigenvalues for heavy and light quark masses ($M_H = 500$(GeV), $M_L = M_b = 5$(GeV)) while giving rise to the different size of FCNC. \begin{enumerate} \item \( M_d = \left[ \begin{array}{cc} 5 & 0 \\ 50 & 500 \end{array} \right] (GeV) \: , \: Z^{bb^\prime} = 10^{-3} \: , \) \item \( M_d = \left[ \begin{array}{cc} 50 & 0 \\ 500 & 50 \end{array} \right] (GeV) \: , \: Z^{bb^\prime} = 10^{-1} \: . \) \end{enumerate} This exercise tells us that the FCNC between singlet quark and ordinary quark is enhanced when the off-diagonal element in the mass matrix is larger than the diagonal element. This kind of analysis can be extended to the general case with arbitrary numbers of isosinglet up-type quarks and down-type quarks. Here we just record the full lagrangian for the model (ref.\cite{handoko}). \begin{eqnarray} {\cal L} & = & {\cal L}_{W^{\pm}} + {\cal L}_{\chi^{\pm}} + {\cal L}_A + {\cal L}_Z + {\cal L}_H + {\cal L}_{\chi^0} , \label{eqn:lagrangian} \end{eqnarray} where, \begin{eqnarray} {\cal L}_{W^{\pm}} & = & \frac{g}{\sqrt{2}} V_{CKM}^{\alpha\beta} \bar{u}^{\alpha} \gamma^{\mu} \, L \, d^{\beta} \, W_{\mu}^+ + h.c. , \\ {\cal L}_{\chi^{\pm}} & = & \frac{g}{\sqrt{2} M_W} V_{CKM}^{\alpha\beta} \bar{u}^{\alpha} \left( m_{u^{\alpha}} L - m_{d^{\beta}} R \right) d^{\beta} \chi^+ + h.c. , \\ {\cal L}_A & = & \frac{e}{3} \left( 2 \bar{u}^{\alpha} \gamma^{\mu} u^{\alpha} - \bar{d}^{\alpha} \gamma^{\mu} d^{\alpha} \right) \, A_{\mu} , \\ {\cal L}_Z & = & \frac{g}{2\cos\theta_W} \left\{ \bar{u}^{\alpha} \gamma^{\mu} \left[\left( {Z_u}^{\alpha\beta} - \frac{4}{3} \sin^2\theta_W \delta^{\alpha\beta} \right) L - \frac{4}{3} \sin^2\theta_W \delta^{\alpha\beta} R \right] u^{\beta} \right. \nonumber \\ & & + \left. \bar{d}^{\alpha} \gamma^{\mu} \left[\left( \frac{2}{3} \sin^2\theta_W \delta^{\alpha\beta} - {Z_d}^{\alpha\beta}\right) L + \frac{2}{3}\sin^2\theta_W \delta^{\alpha\beta} R\right] d^{\beta}\right\} \, Z_{\mu} , \\ {\cal L}_H & = & \frac{-g}{2 M_W} \left[ {Z_u}^{\alpha\beta} \bar{u}^{\alpha} \left(m_{u^{\alpha}} L + m_{u^{\beta}} R\right) u^{\beta} \right. \nonumber \\ & & \left. \: \: \: \: \: \: \: \: + {Z_d}^{\alpha\beta} \bar{d}^{\alpha} \left(m_{d^{\alpha}} L + m_{d^{\beta}} R\right) d^{\beta} \right] \, H , \\ {\cal L}_{\chi^0} & = & \frac{-ig}{2 M_W} \left[ {Z_u}^{\alpha\beta} \bar{u}^{\alpha} \left(m_{u^{\alpha}} L - m_{u^{\beta}} R\right) u^{\beta} \right. \nonumber \\ & & \left. \: \: \: \: \: \: \: \: - {Z_d}^{\alpha\beta} \bar{d}^{\alpha} \left(m_{d^{\alpha}} L - m_{d^{\beta}} R\right) d^{\beta} \right] \, \chi^0 . \end{eqnarray} For $N_u = 3$ and $N_d = 4$, the diagonal elements and off-diagonal elements of FCNC are given by the following equations : \begin{eqnarray} Z^{bb} & = & 1 - \left| {V_L}^{4b} \right|^2 \: , \\ Z^{bs} & = & -{{V_L}^{4b}}^\ast V_L^{4s} \: , \\ Z^{bb^\prime} & = & -{{V_L}^{4b}}^\ast V_L^{44} \: , \end{eqnarray} where $V_L$ is a unitary matrix which diagonalizes the down type quark mass matrix \begin{equation} {d_L}^{0a} = {V_L}^{ab} {d_L}^b \: . \end{equation} In the basis in which up-type quark is diagonalized, $3$ by $4$ part of $V_L$ is just CKM matrix, \begin{equation} {V_L}^{i a} = V_{CKM}^{ia} \: \: (i = 1,2,3 \: , \: a = 1,2,3,4) \: . \end{equation} This leads to the relation between the FCNC and the deviation of unitarity of CKM matrix \begin{equation} Z^{bs} = -{{V_L}^{4b}}^\ast V_L^{4s} = \sum_{i=1}^{3} {V_{CKM}^{ib}}^\ast V_{CKM}^{is} \: . \end{equation} Therefore the unitarity of the CKM matrix no longer holds. This non-unitarity ``quadrangle'' relation is used to constrain the FCNC coupling, e.g. $Z^{bs}$, in later section. However, the deviation is suppressed when the diagonal element of singlet quark mass is infinite and keeping off-diagonal element finite. \subsection{Rare decays and Effect of FCNC} \subsubsection{\it New Physics v.s. Standard Model in $b \rightarrow s Z \rightarrow s l^+ l^-$} \hspace*{1em} In the present beyond standard model, there is FCNC which leads to tree level coupling $Z^{bs}$, $Z^{bd}$ and $Z^{sd}$. In the standard model, the same vertex comes from top quark one loop diagrams. Therefore the condition that the $Z$ FCNC dominates over the standard model contribution in $Z$ sector is : \begin{equation} \left| {Z^{bs} \over {V_{CKM}^{tb}}^\ast V_{CKM}^{ts}} \right| > O(\alpha) \simeq 0.012 \: . \end{equation} Then, even tiny coupling for $Z$ FCNC, it can easily dominates over the standard model contribution. Experimental constraints and quadrangle constraints for the FCNC in $bs$, $bd$ and $sd$ sectors are given in Table (\ref{tab:bound}). {}From Table (\ref{tab:bound}) (refs. \cite{nir}, \cite{data} and \cite{handoko}), $\left|{Z^{bs}}/{{V_{CKM}^{tb}}^\ast V_{CKM}^{ts}} \right| \le 0.05 $ and $ \left| {Z^{bd}}/ {{V_{CKM}^{tb}}^\ast V_{CKM}^{td}} \right| \le 0.79$. Therefore, there are allowed regions for $Z^{bs}$ and $Z^{bd}$ couplings where the tree level $Z$ FCNC can dominate over the 1-loop top quark diagrams in the standard model. We can expect the drastic change of the differential decay rates in the present beyond standard model in $b \rightarrow s l^+ l^-$ and $b \rightarrow d l^+ l^-$ process. \begin{table}[t] \begin{center} \begin{tabular}{|l|l|l|} \hline Decay process & Experiment constraint & Quadrangle constraint \\ \hline \hline $B \rightarrow X_s \mu^+ \mu^-$ & $\left| \frac{Z^{bs}}{{V_{CKM}^{cb}}^\ast V_{CKM}^{cs}}\right| \le 0.047$ & $\left| \frac{{V_{CKM}^{tb}}^\ast V_{CKM}^{ts}} {{V_{CKM}^{cb}}^\ast V_{CKM}^{cs}}\right| \ge 0.94$ \\ \hline $B \rightarrow X_d \mu^+ \mu^-$ & $\left| \frac{Z^{bd}}{{V_{CKM}^{cb}}^\ast V_{CKM}^{cd}}\right| \le 0.23$ & $\left| \frac{{V_{CKM}^{tb}}^\ast V_{CKM}^{td}} {{V_{CKM}^{cb}}^\ast V_{CKM}^{cd}}\right| \ge 0.29$ \\ \hline $K^+ \rightarrow \pi^+ \nu \bar{\nu}$ & $\left| \frac{Z^{sd}}{{V_{CKM}^{cd}}^\ast V_{CKM}^{cs}}\right| \le 2.9 \times 10^{-4}$ & $\left| \frac{{V_{CKM}^{td}}^\ast V_{CKM}^{ts}} {{V_{CKM}^{cd}}^\ast V_{CKM}^{cs}}\right| \ge 0$ \\ \hline \end{tabular} \caption{Upper bound for $\left| Z^{\alpha\beta} \right| $ from experiments, with assuming $Z$ exchange tree diagrams are dominant.} \label{tab:bound} \end{center} \end{table} \subsubsection{\it New Physics v.s. Standard Model in $b \rightarrow s \gamma$ and $b \rightarrow s g$ process} \hspace*{1em} There is no flavor changing neutral currents for $b \rightarrow s \gamma $ and $ b\rightarrow s g $ ($g$ : gluon) processes in the beyond standard model and standard model. Therefore, in order that the new physics contribution dominates over the standard model contribution, the FCNC coupling constant must be the same order as that of the CKM matrix elements. The condition for the new physics contribution dominates over the standard model contribution is : \begin{equation} \left| {Z^{bs} \over {V_{CKM}^{tb}}^\ast V_{CKM}^{ts}} \right| > O(1) \: . \end{equation} {}From Table (\ref{tab:bound}) (refs. \cite{nir}, \cite{handoko}) we can see that there will be no significant contribution to $b \rightarrow s \gamma (g)$ process because $\left| {Z^{bs}}/{{V_{CKM}^{tb}}^\ast V_{CKM}^{ts}} \right| < 0.05$. For $b \rightarrow d \gamma (g)$, new physics contribution can be significant because $\left| {Z^{bd}}/{{V_{CKM}^{tb}}^\ast V_{CKM}^{td}} \right| < 0.79$. Let us summarize the computation of $b \rightarrow d \gamma$ briefly. The amplitude is proportional to the magnetic moment interaction (refs. \cite{inami}, \cite{gautam} and \cite{handoko}), \begin{equation} T = \frac{G_F e}{4 \sqrt{2} \pi^2} \; \bar{d_L} \sigma_{\mu \nu} b_R \epsilon^{\mu} q^{\nu} F \: , \end{equation} where, \begin{eqnarray} F & = & Q_u {V_{CKM}^{tb}}^\ast V_{CKM}^{td} F^{cc} \left( {m_t \over M_W} \right) \nonumber \\ & + & Q_d Z^{db} \left\{ {2 \over 3} \sin^2\theta_W {F_1}^{NC}(0) + 2 {F_3}^{NC}(0) \right\} \\ & + & Q_d Z^{db^\prime} Z^{b^\prime b} \left\{ {F_3}^{NC}\left({m_{b^\prime} \over M_Z}\right) + {F_2}^{NC}\left({m_{b^\prime} \over m_H}, {m_{b^\prime} \over M_Z}\right) \right\} \nonumber \end{eqnarray} where $Q_u = 2/3$ and $Q_d=-1/3$, and $m_H$ is the neutral Higgs mass. The first term comes from the top quark loop. The second and the third terms come from down type quark loop. The second term represents the light down type quarks $(d,b)$ loop and the third term represents the heavy down type quark $(b^\prime)$. The details of the functions and their behavior are given in ref. \cite{handoko}. \subsection{\bf Summary} \hspace*{1em} We study the rare $B$ decays in the non-sequential extention of fermion sector. The effect of vector like down type singlet quark is studied. It has been shown that the tree level FCNC can contribute to $ b \rightarrow s(d) l^+ l^-$ processes, while $b \rightarrow s \gamma (g)$ does not receive the significant contribution due to the New Physics. In $b \rightarrow d \gamma (g)$, the effect of FCNC ($Z^{bd}$, $Z^{db^\prime}$, $Z^{bb^\prime}$) may be seen. \bigskip \noindent {\large \bf Acknowledgments} T. M. would like to thank I. Bigi, J. Kodaira and T. Kouno for comments. This work is supported by the Grant-in-Aid for Scientific Research ($\sharp 06740220$) from the Ministry of Education, Science and Culture, Japan. The work of L.T.H. was supported by a grant from the Overseas Fellowship Program (OFP-BPPT), Indonesia.
\section{Introduction} The relationship between the entropy of a physical black hole and its internal degrees of freedom remains a subject of active research. A natural question to ask in this regard is: can these degrees of freedom be effectively included in a functional integral description of black hole entropy? In an attempt to give an affirmative answer to this question, we investigate in this paper a microcanonical functional integral when applied to a quantum self-gravitating statistical system that includes spacetimes whose topology and boundary conditions coincide with the ones of (either distorted or Kerr-Newman) eternal black holes. A proposal for the density of states of a gravitational system defined in terms of a microcanonical functional integral has been suggested recently in Ref. \cite{BrYo2}. This integral is defined as a formal sum over Lorentzian geometries. The black hole density of states is obtained from this functional integral when the latter is approximated semiclassically by using a complex metric whose boundary data at its single boundary surface coincide with the boundary data of a Lorentzian, stationary black hole. The density of states defined accordingly equals the exponential of one fourth of the area of the black hole horizon. This proposal opens the possibility of determining the thermodynamical properties of black hole systems starting from a sum over real Lorentzian geometries. However, several problems remain in this approach. First, a spacelike hypersurface $\Sigma$ that describes the initial data of a Lorentzian black hole has to cross necessarily the event horizon and eventually intersect the interior singularity. Second, a Lorentzian, stationary black hole is not a extremum of the microcanonical action for a spacetime region with a single timelike boundary surface. This implies that the black hole density of states whose boundary data correspond to the ones of a Lorentzian, stationary black hole cannot be approximated semiclassically by using the same Lorentzian metric that motivates its boundary conditions. It is therefore necessary to use complex metrics to evaluate the Lorentzian functional integral in a steepest descents approximation. This procedure yields the correct result for the entropy but conceals its origin: the interior of the Lorentzian black hole literally disappears by virtue of this procedure, leaving effectively only a periodically identified Euclidean version of the ``right" wedge region of a Kruskal diagram. The properties of the black hole interior become encoded in a set of conditions at the so-called ``bolt" of the complex geometry. In this approach the statistical origin of entropy and its relationship to the internal degrees of freedom of a black hole remain obscure. The problems mentioned above and the role of internal degrees of freedom in functional integral descriptions of black hole thermodynamics can be addressed by explicitly considering the eternal version of a black hole \cite{Ma}. The excitations of the physical black hole can be associated with the deformations of an initial global Cauchy surface $\Sigma$ of the eternal black hole. In general, the spatial slices $\Sigma$ that foliate an eternal black hole are (deformed) Einstein-Rosen bridges with wormhole topology $R^1 \times S^2$. The spacetime is composed of two wedges $M_+$ and $M_-$ located in the right and left sectors of a Kruskal diagram. Internal and external degrees of freedom of the black hole can be easily identified in this approach since the hypersurfaces $\Sigma$ are naturally divided in two parts $\Sigma_+$ and $\Sigma_-$ by a bifurcation two-surface $S_0$ where the lapse function $N$ vanishes. While the ``external" degrees of freedom of the original black hole are naturally given by the initial data at $\Sigma_+$, its ``internal" degrees of freedom can be identified with initial data defined at $\Sigma_-$ \cite{FrMa2}. \section{Microcanonical action and functional integral} Consider a spacetime region whose three-dimensional timelike boundary surface $B$ consists of two disconnected parts $B_+$ and $B_-$. The microcanonical action $S_m$ is the action appropriate to a variational principle in which the fixed boundary conditions at the timelike boundaries $B_\pm$ are not the spacetime three-geometry but the surface energy density $\varepsilon$, surface momentum density $j_a$, and boundary two-metric $\sigma_{ab}$ \cite{BrYo2,BrMaYo}. The covariant form of the microcanonical action for a general spacetime whose timelike surfaces $B_+$ and $B_-$ are located in the regions $M_+$ and $M_-$ respectively has been presented in Ref. \cite{Ma}. Its Hamiltonian form is easily obtained under a $3+1$ spacetime split by recognizing that there exists a direction of time at the boundaries $B_\pm$ \cite{Ma}. By introducing the momentum $P^{ij}$ conjugate to the three-metric $h_{ij}$ of $\Sigma$ for the so-called ``tilted" foliation \cite{FrMa2} and integrating the kinetic part of the volume integral, the action becomes $S_m = \int_M d^4x [ P^{ij} \dot h_{ij} - N{\cal H} - V^i {{\cal H}_i}]$, where the dot denotes differentiation with respect to time, $V^i$ denotes the shift vector, and the gravitational contributions to the Hamiltonian and momentum constraints are given by the usual expressions. Observe that the action vanishes identically for stationary solutions of the vacuum Einstein equations describing stationary eternal black holes. In this case $\dot h_{ij}=0$, the constraint equations are satisfied, and no boundary terms remain in the action. Consider now the microcanonical functional integral for a gravitational system whose timelike boundary surfaces $B_\pm$ are located in $M_\pm$. The path integral \begin{equation} {\bar {\nu}}[\varepsilon_+, j_+, \sigma_+ ; \varepsilon_-, j_-, \sigma_-] = \sum_{\sl M} \int {\cal D}H \exp(i S_m) \label{ournu} \end{equation} is a functional of the energy density $\varepsilon_\pm$, momentum density $j_\pm$, and two-metric $\sigma_{\pm}$ at the boundaries $B_+$ and $B_-$. The sum over $M$ refers to a sum over manifolds of different topologies whose boundaries have topologies $B_+ \equiv S_+ \times S^1 = S^2 \times S^1$ and $B_- \equiv S_- \times S^1 = S^2 \times S^1$. The element $S^1$ is due to the periodic identification in the global time direction at the boundaries when the initial and final hypersurfaces are identified. The integral is a sum over periodic Lorentzian metrics that satisfy the boundary conditions at $B_+$ and $B_-$. The eternal black hole functional integral ${\bar {\nu}}_{*}$ is obtained when the boundary data $(\varepsilon_+, j_+, \sigma_+)$ and $(\varepsilon_-, j_-, \sigma_-)$ of the geometries summed over correspond to the boundary data of a general Lorentzian, stationary eternal black hole. The boundary data of this solution can be determined at $S_+$ and $S_-$ for each slice $\Sigma$. Observe that by virtue of the gravitational constraint equations, these data determine uniquely the size of the black hole horizon \cite{ensembles} and are such that the two-metric is continuous at this horizon. We evaluate now the functional integral in the semiclassical approximation. This requires finding a four-metric that extremizes the action $S_m$ and satisfies the boundary conditions $(\varepsilon_+, j_+, \sigma_+)$ at $S_+$ and $(\varepsilon_-, j_-, \sigma_-)$ at $S_-$. Since the classical Lorentzian eternal black hole metric can be periodically identified and placed on a manifold whose two spatial boundaries have the desired topologies, the resulting metric can be used to approximate the path integral. The periodic identification alters neither the constraint equations nor the boundary data. In the semiclassical approximation the functional integral ${\bar{\nu}}_{*}$ becomes \cite{Ma} \begin{equation} {\bar{\nu}}_{*}[\varepsilon_+,j_+, \sigma_+; \varepsilon_- ,j_-,\sigma_-] \approx \exp \big( i S_m[{\tilde N}, {\tilde V}, {\tilde h}] \big) \approx \exp \big( 0 \big)\ , \label{seminuL} \end{equation} since the microcanonical action $S_m[{\tilde N}, {\tilde V}, {\tilde h}]$ evaluated at the periodically identified geometry vanishes identically if the stationarity condition and the constraints are satisfied. It is illustrative to consider now a complex four-metric which also extremizes the microcanonical action for eternal black hole boundary conditions and which can be used to reevaluate the path integral (\ref{ournu}) in a steepest descent approximation. This alternative approximation of the quantity ${\bar{\nu}}_{*}$ is useful in understanding the relationship of the result (\ref{seminuL}) with the density of states for an ordinary (that is, non-eternal) black hole \cite{BrYo2}. The complex metric can be obtained from the Lorentzian eternal black hole metric by a complexification map $\Psi$ defined by $\Psi(N)= -iN$, $\Psi(V^i) = -iV^i$. This map preserves the reflection symmetry and the canonical variables $h_{ij}$ and $P^{ij}$ of the Lorentzian solution. The complex geometry consists of two complex sectors ${\bar M}_+$ and ${\bar M}_-$ which join at the locus of points at which the lapse vanishes. This geometry is also a solution of Einstein equations if one requires that conical singularities do not exist at that locus for every $\Sigma$. To do this, it is necessary to puncture each complex sector and to close smoothly the geometry at the inner boundaries $^3\! H_\pm = ^2\! H_\pm \times S^1$ of ${\bar M}_\pm$, where $^2\! H_{\pm}$ denotes the intersection of the slices $\Sigma_{\pm}$ with the black hole horizon for the Lorentzian metric \cite{Ma}. After imposing these regularity conditions, the topology of each sector ${\bar M}_\pm$ becomes $R^2 \times S^2$. However, each element $^3\! H_+$ and $^3\! H_-$ does contribute a term to the microcanonical action for the complex geometry. The regular complex metric is not included in the sum over Lorentzian geometries ${\bar{\nu}}_{*}$ in (\ref{ournu}) but can be used to approximate it by distorting the contours of integration for both lapse and shift into the complex plane \cite{BrYo2}. In this approximation the path integral becomes ${\bar{\nu}}_{*} \approx \exp ( i S_m[-i{\tilde N}, -i{\tilde V}, {\tilde h}])$, where $S_m[-i{\tilde N}, -i{\tilde V}, {\tilde h}]$ is the action of the complex metic when the smoothness of the geometries at $^3 \! H_+$ and $^3 \! H_-$ is inforced. This action turns out to be $ S_m[-i{\tilde N}, -i{\tilde V}, {\tilde h}] = -{i} A_+/4 +{i} A_- /4$, where $A_+$ and $A_-$ denote the surface area of the elements ${^2\!H_{\scriptscriptstyle +}}$ and ${^2\! H_{\scriptscriptstyle -}}$ \cite{Ma}. Since the periodic identification and the complexification $\Psi$ do not alter the boundary data nor the gravitational constraint equations, the area $A_+$ of ${^2\!H_{\scriptscriptstyle +}}$ coincides with the area $A_-$ of ${^2\!H_{\scriptscriptstyle -}}$: $A_+ (\varepsilon_+, j_+, \sigma_+) = A_- (\varepsilon_-, j_-, \sigma_-) \equiv A_H$. This implies that, in agreement with (\ref{seminuL}), the eternal black hole functional integral is ${\bar{\nu}}_{*}[\varepsilon_+ ,j_+, \sigma_+; \varepsilon_-, j_-, \sigma_-] \approx \exp{( A_H/4 - A_H/4)} = \exp (0)$ in the ``zero-loop" approximation. If the microcanonical functional integral (\ref{ournu}) is interpreted as the density of states of the statistical system, it is possible to express ${\bar{\nu}}_{*}$ approximately as ${\bar{\nu}}_{*}[\varepsilon_+, j_+, \sigma_+; \varepsilon_-, j_-, \sigma_-] \approx \exp({\cal S} [\varepsilon_+, j_+, \sigma_+; \varepsilon_-, j_-, \sigma_-]) $, where ${\cal S}$ represents the total entropy of the system. The above result implies that the entropy for the system in the semiclassical approximation is \begin{equation} {\cal S} \approx {1\over 4}{A_H} - {1\over 4} {A_H} = 0 \ , \label{entropy} \end{equation} where $A_H$ is the area of the horizon of the physical eternal black hole solution that classically approximates the system \cite{Ma}. The total entropy is given formally by the subtraction ${\cal S} = {{\cal S}_+}[\varepsilon_+ ,j_+, \sigma_+] - {{\cal S}_-}[\varepsilon_- ,j_-, \sigma_-]$, where ${{\cal S}_+}$ and ${{\cal S}_-}$ can be interpreted as the semiclassical entropies associated with the external ($M_+$) and internal ($M_-$) regions respectively of the eternal black hole system. \section{Conclusions} The functional integral (\ref{seminuL}) refers to a quantum-statistical system which is classically approximated by a general stationary, eternal black hole solution of Einstein equations within a region bounded by two timelike surfaces $B_+$ and $B_-$. Its semiclassical value is a consequence of the choice of boundary data, the gravitational constraint equations, and the vanishing of the microcanonical action for the four-geometries that satisfy the boundary conditions and approximate the path integral. The calculation presented above applies to any distorted black hole in the strong gravity regime. It indicates that a pure state (of zero entropy) can be defined not only for matter fields perturbations propagating in the spacetime of an eternal black hole but also for the gravitational field itself. This is physically appealing: the initial data for the eternal black hole specified at the spacelike hypersurface $\Sigma$ contain all the information required for the evolution of both the exterior and interior parts of a physical black hole. The entropy associated with $\Sigma$ must therefore equal zero. Since in a microcanonical description it seems natural to relate the external and internal degrees of freedom of a black hole with the boundary data at the surfaces $B_+$ and $B_-$ respectively \cite{FrMa2}, we believe that the microcanonical functional integral for eternal black hole systems opens the possibility of extending path integral formulation of black hole thermodynamics to situations when internal degrees of freedom are present and allows the study of gravitational statistical properties in terms of a single pure state \cite{Ma}. These conclusions are in complete agreement with thermofield dynamics descriptions of quantum processes and, in particular, with the application of this approach to black hole thermodynamics developed originally by Israel \cite{Is} for small perturbations. They strongly suggest that the thermofield dynamics description of quantum field processes in a curved background can be extended beyond perturbations to the gravitational field itself of distorted eternal black holes. I am grateful to Valeri Frolov and Werner Israel for helpful comments. This work was supported by the Natural Sciences and Engineering Research Council of Canada.
\section{Introduction} In the Minimal Supersymmetric Standard Model (MSSM) \cite{1,2} two Higgs doublets are necessary, leading to five physical Higgs bosons $h^0, H^0, A^0$, and $H^\pm$ \cite{3,4}. If all supersymmetric (SUSY) particles are very heavy, the charged Higgs boson $H^+$ decays dominantly into $t \bar b$; the decays $H^+ \rightarrow \tau^+ \nu$ and/or $H^+ \rightarrow W^+ h^0$ are dominant below the $t \bar b$ threshold \cite{3,4b}. In ref.~\cite{5} all decay modes of $H^+$ including the SUSY-particle modes were studied in detail; it was shown that the SUSY decay modes $H^+ \rightarrow \ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i \bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j (i,j = 1,2)$ can be dominant in a large region of the MSSM parameter space due to large $t$ and $b$ quark Yukawa couplings and large $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi$- and $\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi$-mixings, and that this could have a decisive impact on $H^+$ searches at future colliders. Here $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i \, (\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j)$ are the scalar top (scalar bottom) mass eigenstates which are mixtures of $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_L$ and $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_R$ ($\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_L$ and $\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_R$). The standard QCD corrections are very large for the width of $H^+ \rightarrow c \bar s$ and can be large ($+10$\% to $-50$\%) for that of $H^+\rightarrow t \bar b$ \cite{6a}. The QCD corrections from the SUSY-particle loops are calculated within the MSSM for $H^+\rightarrow t \bar b$ in \cite{6b} and turn out to be non-negligible ($\sim$ 10\%) for certain values of the MSSM parameters. This suggests that the QCD corrections to $H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}$ could also be large. Therefore it should be examined whether the result in \cite{5} remains valid after including the QCD corrections. In this paper we calculate the ${\cal O}(\alpha_s)$ QCD corrections to the width of $H^+ \rightarrow \ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i \bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j$ within the MSSM. To the best of our knowledge they are not known in the literature. We obtain the complete ${\cal O}(\alpha_s)$ corrected width in the \ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi renormalization scheme (i.e. the \ifmmode{\overline{\rm MS}} \else{$\overline{\rm MS}$} \fi scheme with dimensional reduction \cite{7}) including all quark mass terms and $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_L-\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_R$ mixings. The main complication here is that the $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_L-\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_R$ mixing angles are renormalized by the SUSY QCD corrections. We find that the corrections to the $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}$ width are significant but that the $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}$ mode is still dominant in a wide parameter range. \section{Tree level result} We first review the tree level results \cite{5}. The squark mass matrix in the basis ($\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_L$, $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_R$), with $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi=\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi$ or $\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi$, is given by \cite{3,4} \begin{equation} \left( \begin{array}{cc}m_{LL}^2 & m_{LR}^2 \\ m_{RL}^2 & m_{RR}^2 \end{array} \right)= (R^{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi})^{\dagger}\left( \begin{array}{cc}m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_1}^2 & 0 \\ 0 & m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_2}^2 \end{array} \right)R^{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}, \end{equation} where \begin{eqnarray} m_{LL}^2 &=& M_{\tilde{Q}}^2+m_q^2+m_Z^2\cos 2\beta (I_q-Q_q\sin^2\theta_W), \\ m_{RR}^2 &=& M_{\{\tilde{U}, \tilde{D}\}}^2+m_q^2+m_Z^2\cos 2\beta Q_q\sin^2 \theta_W, \\ m_{LR}^2=m_{RL}^2 &=& \left\{ \begin{array}{ll} m_t(A_t-\mu\cot\beta) & (\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi=\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi) \\ m_b(A_b-\mu\tan\beta) & (\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi=\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi) \end{array} \right. , \end{eqnarray} and \begin{equation} \label{5} R^{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}_{i\alpha}=\left( \begin{array}{cc}\cos\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi & \sin\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi \\ -\sin\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi & \cos\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi \end{array}\right) . \end{equation} Here the mass eigenstates $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i(i=1,2)$ (with $m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_1}<m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_2}$) are related to the SU(2)$_L$ eigenstates $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_{\alpha}(\alpha=L,R)$ as $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i=R^{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}_{i\alpha}\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_{\alpha}$. Note that in the sign convention used here the parameters $A_{t,b}$ correspond to $(-A_{t,b})$ of ref.\cite{5}. The tree-level decay width of $H^+ \rightarrow \ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i \bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j$ is then given by (see Fig.~1a) \begin{equation} \label{6} \Gamma^{(0)}(H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j) =\frac{N_C\kappa}{16\pi m_H^3}|G_{ij}|^2\,, \end{equation} where $m_H$ is the $H^+$ mass, $\kappa=\kappa(m_H^2, m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2, m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2 )$, $\kappa(x,y,z)\equiv ((x-y-z)^2-4yz)^{1/2}$, $N_C=3$, and \begin{eqnarray} \label{7} G_{ij}&=& \frac{g}{\sqrt{2}m_W} R^{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi}\left( \begin{array}{cc} m_b^2\tan\beta+m_t^2\cot\beta-m_W^2\sin 2\beta & m_b(A_b\tan\beta+\mu) \\ m_t(A_t\cot\beta+\mu) & 2m_tm_b/\sin 2\beta \end{array} \right)(R^{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi})^{\dagger} \nonumber\\ \end{eqnarray} are the $H^+\bar{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi}_i\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j$ couplings \cite{3,4}, with $g$ being the SU(2) coupling. \section{QCD virtual corrections} The ${\cal O}(\alpha_s)$ QCD virtual corrections to $H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j$ stem from the diagrams of Fig.~1b (vertex corrections) and 1c (wave function corrections). For simplicity we use in this paper the \ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi renormalization scheme\footnote{Strictly speaking, our renormalization scheme is the $\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi'$ scheme \cite{7b} where the ``$\epsilon$-scalar mass'' is absorbed into $M_{\tilde{Q},\tilde{U},\tilde{D}}^2$.} for all parameters which receive the QCD corrections, i.e. $m_{t,b}$, $A_{t,b}$, and $M_{\tilde{Q},\tilde{U},\tilde{D}}$. The renormalized squark mixing angle $\theta_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}$ is then defined by the relations (1--5) in terms of the \ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi parameters $m_{t,b}$, $A_{t,b}$, and $M_{\tilde{Q},\tilde{U},\tilde{D}}$. The one-loop corrected decay amplitudes $G_{ij}^{\rm corr}$ are expressed as \begin{equation} G_{ij}^{\rm corr}=G_{ij}+\delta G_{ij}^{(v)}+\delta G_{ij}^{(w)}, \end{equation} where $G_{ij}$ are defined by (7) in terms of the \ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi parameters, and $\delta G_{ij}^{(v)}$ and $\delta G_{ij}^{(w)}$ are the vertex and squark wave function corrections, respectively. The vertex corrections $\delta G_{ij}^{(v)}$ are calculated from the graphs of Fig.~1b as \begin{eqnarray} \delta G_{ij}^{(v)}&=& \frac{\alpha_s C_F}{4\pi} \left[ \{B_0(m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2, 0, m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2)+B_0 (m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2, 0, m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2) -B_0(m_H^2, m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2, m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2) \right. \nonumber \\ && -2(m_H^2-m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2-m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2) C_0(m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2, \lambda^2, m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2)\}G_{ij} \nonumber \\ && -B_0(m_H^2, m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_k}^2, m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_l}^2)G_{kl}S^{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi}_{ik}S^{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_{lj} \nonumber \\ && +2\{ (\alpha_{LL})_{ij}(m_ty_2+m_by_1)+(\alpha_{RR})_{ij}(m_ty_1+m_by_2)\} B_0(m_H^2, m_b^2, m_t^2) \nonumber\\ && +2\{ (\alpha_{LL})_{ij}m_ty_2+(\alpha_{LR})_{ij}m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}y_1 +(\alpha_{RL})_{ij}m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}y_2+ (\alpha_{RR})_{ij}m_ty_1\} B_0(m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_t^2) \nonumber\\ && + 2\{ (\alpha_{LL})_{ij}m_by_1+(\alpha_{LR})_{ij}m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}y_1 +(\alpha_{RL})_{ij}m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}y_2+ (\alpha_{RR})_{ij}m_by_2\} B_0(m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_b^2) \nonumber\\ && +2\{ (m_t^2+m_b^2-m_H^2)m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi} ((\alpha_{LR})_{ij}y_1+(\alpha_{RL})_{ij}y_2) \nonumber\\ && +(m_b^2+m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2-m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2)m_t ((\alpha_{LL})_{ij}y_2+(\alpha_{RR})_{ij}y_1) \nonumber\\ && +(m_t^2+m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2-m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2)m_b ((\alpha_{LL})_{ij}y_1+(\alpha_{RR})_{ij}y_2) \nonumber\\ && \left. + 2m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}m_tm_b((\alpha_{LR})_{ij}y_2+(\alpha_{RL})_{ij}y_1) \} C_0(m_b^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_t^2) \right] , \end{eqnarray} where $C_F=4/3$, \begin{equation} S^{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}=\left( \begin{array}{cc} \cos 2\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi & -\sin 2\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi \\ -\sin 2\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi & -\cos 2\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi \end{array} \right) , \end{equation} \[ \alpha_{LL}=\left( \begin{array}{rr} \cos\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\cos\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi & -\cos\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\sin\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi \\ -\sin\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\cos\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi & \sin\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\sin\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi \end{array} \right) , \;\;\; \alpha_{LR}=\left( \begin{array}{rr} -\cos\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\sin\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi & -\cos\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\cos\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi \\ \sin\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\sin\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi & \sin\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\cos\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi \end{array} \right) , \] \begin{equation} \alpha_{RL}=\left( \begin{array}{rr} -\sin\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\cos\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi & \sin\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\sin\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi \\ -\cos\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\cos\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi & \cos\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\sin\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi \end{array} \right) , \;\;\; \alpha_{RR}=\left( \begin{array}{rr} \sin\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\sin\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi & \sin\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\cos\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi \\ \cos\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\sin\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi & \cos\theta_\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\cos\theta_\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi \end{array} \right) , \end{equation} \begin{equation} y_1=\frac{g}{\sqrt{2}m_W}m_b\tan\beta=h_b\sin\beta, \;\;\; y_2=\frac{g}{\sqrt{2}m_W} m_t\cot\beta=h_t\cos\beta, \end{equation} and $m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}$ is the gluino mass. A gluon mass $\lambda$ is introduced to regularize the infrared divergences. Here we define the functions $A$, $B_0$, $B_1$ and $C_0$ as in \cite{pv} ($\Delta = 2/(4-D) - \gamma_E +\log 4\pi$): \begin{eqnarray} A(m^2)&=& \int\frac{d^Dq}{i\pi^2}\frac{1}{q^2-m^2}=m^2(\Delta +\log(Q^2/m^2)+1), \\ B_0(k^2, m_1^2, m_2^2)&=& \int\frac{d^Dq}{i\pi^2}\frac{1}{(q^2-m_1^2)((q+k)^2-m_2^2)} \nonumber\\ &=&\Delta-\int_0^1dz\log\frac{(1-z)m_1^2+zm_2^2-z(1-z)k^2-i\delta}{Q^2}, \\ B_1(k^2, m_1^2, m_2^2)&=& \frac{1}{k_{\mu}}\int\frac{d^Dq}{i\pi^2} \frac{q_{\mu}}{(q^2-m_1^2)((q+k)^2-m_2^2)} \nonumber\\ &=&-\frac{\Delta}{2}+\int_0^1dz\,z\log \frac{(1-z)m_1^2+zm_2^2-z(1-z)k^2-i\delta}{Q^2}, \\ C_0(m_1^2, m_2^2, m_3^2)&=& \int\frac{d^Dq}{i\pi^2}\frac{1}{(q^2-m_1^2)((q+k_2)^2-m_2^2) ((q+p)^2-m_3^2)} \nonumber\\ &=& -\int_0^1dx\int_0^1dy\int_0^1dz\delta(1-x-y-z)\, \times \nonumber\\ &&(xm_1^2+ym_2^2+zm_3^2-xym_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2 -yzm_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2-xzm_H^2-i\delta)^{-1}. \end{eqnarray} Here $p$ and $k_2$ are respectively the external momenta of $H^+$ and $\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j$, and $Q$ is the \ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi renormalization scale. Note that $\Delta$ is omitted in the \ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi scheme. The squark wave function corrections $\delta G_{ij}^{(w)}$ are expressed as \begin{equation} \label{17} \delta G_{ij}^{(w)}=-\frac{1}{2}\left[ \dot{\Pi}_{ii}^{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi}(m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2) +\dot{\Pi}_{jj}^{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}(m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2)\right] G_{ij} -\frac{\Pi_{ii'}^{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi}(m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2)}{m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2-m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_{i'}}^2} G_{i'j} -\frac{\Pi_{j'j}^{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}(m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2)}{m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2-m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_{j'}}^2} G_{ij'}, \end{equation} where $i\neq i'$ and $j\neq j'$. $\Pi_{ij}^{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}(k^2)$ are the one-loop corrections to the two-point functions of $\bar{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}_i\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_j$, which are obtained from the graphs of Fig.~1c. $\dot{\Pi}(k^2)$ denotes the derivative with respect to $k^2$. The last two terms in (\ref{17}) represent the corrections due to the renormalization of the $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi$-mixings. The explicit forms are \begin{eqnarray} \dot{\Pi}_{ii}^{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2)&=& \frac{\alpha_sC_F}{4\pi}\left[ -3B_0(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, 0, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2) -2B_1(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, 0, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2) -4m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2\dot{B}_0(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, \lambda^2, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2) \right. \nonumber\\ &&-2m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2\dot{B}_1(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, 0, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2) -4m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2\dot{B}_0(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_q^2) -4B_1(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_q^2) \nonumber\\ &&\left. -4m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2\dot{B}_1(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_q^2) +(-)^{i-1}4\sin 2\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi m_qm_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}\dot{B}_0(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_q^2) \right] , \end{eqnarray} and \begin{equation} \Pi_{i'i}^{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2)=\frac{\alpha_sC_F}{4\pi} \left[ \frac{1}{2}\sin 4\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi (A(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_2}^2)-A(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_1}^2)) +4\cos 2\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi m_qm_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}B_0(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_q^2)\right] . \end{equation} The one-loop corrected decay width in the \ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi scheme is then given by \begin{equation} \Gamma(H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j) =\frac{N_C\kappa_{\rm pole}}{16\pi m_H^3}[|G_{ij}|^2 +2G_{ij}{\rm Re}(\delta G_{ij}^{(v)} +\delta G_{ij}^{(w)})]. \label{20} \end{equation} Here $\kappa_{\rm pole}$ refers to $\kappa$ in (6) evaluated with pole squark masses. The width of (20) is infrared divergent. In the numerical analysis we take the pole quark masses as inputs. The \ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi quark masses are obtained from the pole quark masses by using \begin{eqnarray} m_q(Q)_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi}&=&m_q({\rm pole})-\frac{\alpha_sC_F}{4\pi} \left[ 2m_q(B_0(m_q^2, 0, m_q^2)-B_1(m_q^2, 0, m_q^2)) \right. \nonumber\\ &&+\sin 2\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}(B_0(m_q^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_1}^2) -B_0(m_q^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_2}^2))\nonumber\\ && \left. +m_q(B_1(m_q^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_1}^2) +B_1(m_q^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_2}^2))\right] , \end{eqnarray} which is derived from the graphs of Fig.~1d. Furthermore, in the phase space term $\kappa_{\rm pole}$ in (20) we have to take the pole squark masses given by \begin{eqnarray} m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2({\rm pole})&=&m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2(Q)_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi}-\Pi_{ii}^{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi}(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2) \nonumber \\ &=&m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2(Q)_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi}+\frac{\alpha_sC_F}{4\pi} \left[ -3A(0)+4m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2B_0(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, 0, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2) +2m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2B_1(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, 0, m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2) \right. \nonumber \\ &&-\cos^22\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi A(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2)-\sin^22\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi A(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_{i'}}^2)+4A(m_q^2) \nonumber\\ &&+4m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2B_0(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_q^2) +4m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2B_1(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_q^2) \nonumber\\ &&\left. -(-)^{i-1}4\sin 2\theta_\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi m_qm_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi} B_0(m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}^2, m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}^2, m_q^2) \right] . \label{22} \end{eqnarray} \section{Gluon emission} The infrared divergences in (\ref{20}) are cancelled by including the ${\cal O}(\alpha_s)$ contribution from real gluon emission from $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi$ and $\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}$ (Fig.~1e). The decay width of $H^+(p)\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i(k_1)+\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j(k_2)+g(k_3)$ is given in terms of the \ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi parameters as \begin{equation} \Gamma(H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_jg) =\frac{\alpha_sC_FN_C|G_{ij}|^2}{4\pi^2 m_H} [(m_H^2-m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2-m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2)I_{12} -m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}^2I_{11}-m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}^2I_{22}-I_1-I_2]. \label{23} \end{equation} The functions $I_n$, and $I_{nm}$ are defined as \cite{8} \begin{equation} I_{i_1\ldots i_n}=\frac{1}{\pi^2} \int\frac{d^3k_1}{2E_1}\frac{d^3k_2}{2E_2}\frac{d^3k_3}{2E_3} \delta^4(p-k_1-k_2-k_3)\frac{1} {(2k_3k_{i_1}+\lambda^2)\ldots(2k_3k_{i_n}+\lambda^2)}. \end{equation} The explicit forms of $I_{i_1\ldots i_n}$ are given in \cite{8}. In (\ref{23}), $I_{11,22,12}$ are infrared divergent. We have checked that the infrared divergences in (\ref{23}) cancel those in (\ref{20}). In the numerical analysis we define the corrected decay width as $\Gamma^{\rm corr}(H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j)\equiv \Gamma(H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j)+ \Gamma(H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_jg)$. \section{Numerical results and conclusions} As in ref.\cite{5}, we choose \{ $m_H$, $m_{t,b}$(pole), $M$, $\mu$, $\tan\beta$, $M_{\tilde{Q}}$, $A$ \} as the basic input parameters of the MSSM, taking $M=(\alpha_2/\alpha_s)m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}=(3/5\tan^2\theta_W)M'$, $M_{\tilde{Q}}\equiv M_{\tilde{Q}}(Q)_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi}= M_{\tilde{U}}(Q)_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi}=M_{\tilde{D}}(Q)_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi}=M_{\tilde{L}}$ and $A\equiv A_t(Q)_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi}=A_b(Q)_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi}=A_{\tau}$. Here $M$ ($M'$) is the SU(2) (U(1)) gaugino mass, $\alpha_2=g^2/4\pi$, and ($M_{\tilde{L}}$, $A_{\tau}$) are the mass matrix parameters of the slepton sector \cite{5}. The parameters $M$, $M'$, $M_{\tilde{L}}$, and $A_{\tau}$ do not receive ${\cal O}(\alpha_s)$ QCD corrections. The theoretical and experimental constraints for the basic input parameters are described in ref.\cite{5}. We take $m_Z=91.2$GeV, $m_W=80$GeV, $m_t({\rm pole})=180$GeV \cite{mtop}, $m_b({\rm pole})=5$GeV, $\sin^2\theta_W=0.23$ and $\alpha_2=\alpha_2(m_Z)=\alpha/\sin^2\theta_W =(1/129)/0.23=0.0337$. For the running QCD coupling at the renormalization scale $Q$, $\alpha_s=\alpha_s(Q)$, we always take the one-loop expression $\alpha_s(Q)= 12\pi/\{(33-2n_f)\ln(Q^2/\Lambda_{n_f}^2)\}$, with $\alpha_s(m_Z)=0.12$, and the number of quark flavors $n_f=5(6)$ for $m_b<Q\le m_t$ (for $Q>m_t$). We define the QCD corrections as the difference between the ${\cal O}(\alpha_s)$ corrected width $\Gamma^{\rm corr}_{ij}\equiv \Gamma^{\rm corr}(H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j)$ of eqs.(20) plus (23) and the tree-level width $\Gamma_{ij}^{\rm tree}\equiv \Gamma^{(0)}(H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j)$ of eq.(6) where ($m_q(\rm pole)$, $M_{\tilde{Q}}$, $A$) are substituted for ($m_q$, $M_{\tilde{Q},\tilde{U},\tilde{D}}$, $A_q$). The QCD corrections depend on the renormalization scale $Q$. We choose the optimum value of $Q$ ($Q_{\rm opt}$) such that $\displaystyle\Delta(Q)\equiv\sum_{q=t,b}\sum_{i=1,2} (m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}({\rm tree})-m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}({\rm pole}))^2$ is minimized, where $m_{\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i}(\rm tree)$ refers to the $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_i$ mass defined by (1--4) calculated with $m_q(\rm pole)$, $M_{\tilde{Q}}$, and $A$. In order not to vary too many parameters, in the following we fix $\mu=300$GeV, and take the values of $M$ and $\tan\beta$ such that $m_{\tilde{\chi}_1^0}\simeq 50$GeV as in \cite{5} where $\tilde{\chi}_1^0$ is the lightest neutralino. In Fig.2 we show the $m_H$ dependence of the tree-level and corrected widths $\displaystyle\Gamma^{\rm tree}(\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi})\equiv \sum_{i,j=1,2}\Gamma^{\rm tree}_{ij}$ and $\displaystyle\Gamma^{\rm corr}(\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi})\equiv \sum_{i,j=1,2}\Gamma^{\rm corr}_{ij}$, and the tree-level branching ratio $\displaystyle B^{\rm tree}(\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi})\equiv \sum_{i,j=1,2}B^{\rm tree}(H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j)$ \cite{5} for (a) $M_{\tilde{Q}}=250$GeV, $A=650$GeV, $\tan\beta=2$, $M=120$GeV, and (b) $M_{\tilde{Q}}=136$GeV, $A=260$GeV, $\tan\beta=12$, $M=110$GeV. In these two cases we have (in GeV units): (a) $Q_{\rm opt}=216.5$, $m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}=380$, $(m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_1},m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_2}, m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_1},m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_2})\{(\rm tree),(Q_{\rm opt})_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi}, (\rm pole)\}=$ $\{(60,429,251,254)$, $(70,420,251,254)$, $(58,428,258,262)\}$, $m_{\tilde{\chi}_1^+}=94$, and (b) $Q_{\rm opt}=213.5$, $m_{\ifmmode{\tilde{g}} \else{$\tilde{g}$} \fi}=349$, $(m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_1},m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_2}, m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_1},m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_2})\{(\rm tree), (Q_{\rm opt})_{\ifmmode{\overline{\rm DR}} \else{$\overline{\rm DR}$} \fi},(\rm pole)\}=$ $\{(81,302,62,193)$, $(74,292,120,164)$, $(66,302,126,171)\}$, $m_{\tilde{\chi}_1^+}=98$. Here $\tilde{\chi}_1^+$ is the lighter chargino. In both cases we see that the $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}$ mode dominates the $H^+$ decay in a wide $m_{H^+}$ range at the tree level, and that the QCD corrections to the $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}$ mode are significant, but that as a whole they do not invalidate the $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}$ mode dominance. In Table 1 we show the values of the $B^{\rm tree}(\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi})$, the QCD corrections $C\equiv(\Gamma^{\rm corr}(\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi})- \Gamma^{\rm tree}(\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi})) /\Gamma^{\rm tree}(\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi})$, and $C_{ij}\equiv(\Gamma^{\rm corr}_{ij} -\Gamma^{\rm tree}_{ij})/ \Gamma^{\rm tree}_{ij}$ for typical values of $M_{\tilde{Q}}$ and $A$, for (a) $m_{H^+}=400$GeV, $\tan\beta=2$, $M=120$GeV, and (b) $m_{H^+}=400$GeV, $\tan\beta=12$, $M=110$GeV. We see again that the $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}$ mode dominates in a wide region also when the QCD corrections are included. The QCD corrections can be very large at some points of $(M_{\tilde{Q}},A)$; e.g. $C=-0.692$ at (175GeV, 0GeV) and $C_{12}=0.734$ at (225GeV, $-350$GeV) in Table 1b. This occurs when $m_H\sim m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}+m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}$. This enhancement is just a kinematical effect due to the QCD corrections to $m_{\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i}$ and $m_{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi_j}$. In conclusion, we have calculated the ${\cal O}(\alpha_s)$ QCD corrections to the decay width of $H^+\rightarrow\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi_i\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}_j$, including all quark mass terms and $\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_L-\ifmmode{\tilde{q}} \else{$\tilde{q}$} \fi_R$ mixing. We find that the QCD corrections are significant but that they do not invalidate our previous conclusion at tree-level about the dominance of the $\ifmmode{\tilde{t}} \else{$\tilde{t}$} \fi\bar{\ifmmode{\tilde{b}} \else{$\tilde{b}$} \fi}$ mode in a wide MSSM parameter region. \section*{Acknowledgements} The work of Y.Y. was supported in part by the Fellowships of the Japan Society for the Promotion of Scienceand the Grant-in-Aid for Scientific Research from the Ministry of Education, Science and Culture of Japan, No. 06-1923 and 07-1923. The work of A.B., H.E., and W.M. was supported by the ``Fonds zur F\"orderung der wissenschaftlichen Forschung'' of Austria, project no. P10843-PHY. The authors are grateful to Y. Kizukuri for the collaboration at the early stage of this work. \clearpage
\section{Theoretical Treatment of Charm and \section{Introduction} \subsection{Goals and Obstacles} As explained in detail elsewhere in this review, a precise measurement of the lifetimes of the various weakly decaying charm and beauty hadrons possesses great experimental value per se as well as in searches for $B^0 - \bar B^0$ and $D^0 - \bar D^0$ oscillations. Yet there exists also a strong theoretical interest in determining and interpreting those lifetimes; I want to sketch that first in rather qualitative terms and give specifics later. Weak decays of hadrons depend on fundamental parameters of the Standard Model, in particular on the KM parameters and quark masses. It is eminently important to reliably determine their values from data. Alas -- this is easier said than done theoretically (and experimentally)! For in such an endeavour we have to face the "Dichotomy of the Two Worlds". On the one hand there is the "Theorists' World" where quarks and gluons are the relevant strongly interacting entities; it is in this short-distance or Femto World where theorists like to formulate their fundamental theories. On the other hand there is the "Real World" where hadrons constitute the relevant degrees of freedom; it is in that world where everyone (including theorists) lives and measurements are performed. To formulate predictions from the Theorists' World in the language of the Real World and to translate findings from the Real World back into the idioms of the Theorists' World represents the theoretical challenge one faces. One quantitative measure for the difference between the two worlds is provided by the lifetimes of the weakly decaying hadrons carrying the same flavour. On the quark level there is obviously only a single lifetime for a given flavour. Yet in the `Real World' hadrons carrying the same flavour quantum number possess different (and even vastly different) lifetimes; e.g., $\tau (K^+)/\tau (K_S)\simeq 140$, $\tau (D^+)/\tau (D^0)\simeq 2.5$ and $0.9 \leq \tau (B^-)/\tau (B_d)\leq 1.2$. On the other hand deviations of the lifetime ratios from unity evidently decrease for an increasing heavy-flavour mass. This is as expected in a simple `two-component' picture: in the limit of $m_Q$ -- the mass of the heavy-flavour quark $Q$ -- going to infinity the `Spectator Ansatz' should hold where the lifetimes of all hadrons $H_Q$ containing $Q$ coincide; for finite values of $m_Q$ there are {\em pre-asymptotic} corrections. Two lines of reasoning support this qualitative picture: \noindent ($\alpha$) The decay width of a quark $Q$ increases very quickly with its mass $m_Q$: $$\Gamma _Q \propto G_F^2 m_Q^5 \eqno(1)$$ for $m_Q \ll M_W$ changing into $\Gamma _Q \propto \alpha m_Q^3/M_W^2$ for $m_Q \gg M_W$ \footnote{The asymptotic scaling behaviour $\Gamma _Q \propto m_Q^3$ (rather than $\Gamma _Q\propto m_Q$) reflects the coupling of the {\em longitudinal} $W$ boson -- the original Higgs field -- to the top quark.}. Therefore for $m_Q$ sufficiently large, its decay width is bound to exceed $\Lambda _{QCD}$, i.e. the quark $Q$ decays before it can hadronize \footnote{For top decays this happens for $m_t\geq 130$ GeV \cite{RAPALLO}.}. Then there is obviously a universal lifetime for such a heavy flavour and the spectator ansatz applies trivially. \noindent ($\beta$) Analysing all {\em non}-spectator reactions explicitely, one finds their widths to increase with a smaller power of $m_Q$ than the spectator process; thus one arrives at the spectator ansatz as the asymptotic case. The most relevant phenomenological question then is how quickly the limit of universal lifetimes is approached. \subsection{Phenomenology: Legends with Truths} Some mechanisms had been identified very early on that generate differences in the lifetimes of hadrons $H_Q$ with the same heavy flavour $Q$ : "Weak Annihilation" (=WA) of $Q$ with the light valence antiquark for mesons or "W Scattering" (=WS) with the valence diquark system for baryons \footnote{A distinction is often made between W exchange in the s and in the t channel with the former case referred to as `weak annihilation' and the latter as `W exchange'. This classification is however artificial since the two operators mix already under one-loop renormalization in QCD, as discussed later on. Both cases will summarily be referred to as WA.}. Such an analysis had first been undertaken for charm decays. Since WA contributes to Cabibbo allowed decays of $D^0$, but not of $D^+$ mesons (in the valence quark description), it creates a difference in $\tau (D^0)$ vs. $\tau (D^+)$. However the WA rate is doubly suppressed relative to the spectator rate, namely by the helicity factor $(m_q/m_c)^2$ with $m_q$ denoting the largest mass in the final state and by the `wavefunction overlap' factor $(f_D/m_c)^2$ reflecting the practically zero range of the low-energy weak interactions: $$\Gamma ^{(0)}_{W-X}(D^0) \propto G_F^2f^2_Dm_q^2m_c \, .\eqno(2)$$ Therefore it had originally been suggested that already charm hadrons should possess approximately equal lifetimes. It then came as quite a surprise when observations showed it to be otherwise -- in particular since the first data suggested a considerably larger value for $\tau (D^+)/\tau (D^0)$ than measured today. This near-shock caused a re-appraisal of the theoretical situation; its results at that time can be summarized in four main points: \noindent (i) One source for a lifetime difference had too quickly been discarded as insignificant. Cabibbo-allowed nonleptonic decays of $D^+$ -- but not of $D^0$ -- mesons produce two antiquarks in the final state that carry the same flavour: $$D^+ = [c\bar d] \rightarrow (s\bar d u) \bar d $$ Thus one has to allow for the {\em interference} between different quark diagrams in $D^+$ , yet not in $D^0$ decays; the $\bar d$ valence antiquark in $D^+$ mesons thus ceases to play the role of an uninvolved bystander and a difference in the $D^+$ vs. $D^0$ lifetimes will arise. While this interference had been included in descriptions of the $D\rightarrow K \pi$ two-body modes it was ignored for total widths. For it was thought (often without stating it explicitely) that the required coherence would not be maintained between the two amplitudes when applied to inclusive transitions. This assumption was challenged in ref.\cite{PI} where it was argued that even the two inclusive amplitudes remain sufficiently coherent. The interference turns out to be destructive, i.e. it prolongs $\tau (D^+)$ over $\tau (D^0)$, but only once the QCD radiative corrections have been included. This effect is usually referred to as `Pauli Interference' (=PI) although such a name would be misleading if it is interpreted as suggesting that the interference is automatically destructive. \noindent (ii) It was argued \cite{SONI} that the helicity suppression of the WA contribution to $D$ decays can be vitiated. The diagram in Fig.1 contains gluon bremsstrahlung off the initial antiquark line in the $W$-exchange reaction; evaluating this particular diagram explicitely one finds: $$\Gamma ^{(1)}_{W-X}(D^0)\propto (\alpha_s /\pi ) G_F^2(f_D/\aver{E_{\bar q}})^2m_c^5 \eqno(3)$$ with $\aver {E_{\bar q}}$ denoting the average energy of the initial antiquark $\bar q$ \footnote{The $1/\aver {E_{\bar q}}$ term in the amplitude derives from the propagator in the diagram.}. Using a non-relativistic wavefunction for the decaying meson one has $\aver{E_{\bar q}}\simeq m_q$. This contribution, although of higher order in $\alpha_s$, would dominate over the lowest order term $\Gamma ^{(0)}_{W-X}$ since helicity suppression has apparently been vitiated and the decay constant $f_D$ is now calibrated by $\aver{E_{\bar q}}$ with $f_D/\aver{E_{\bar q}} \sim {\cal O}(1)$ rather than $f_D/m_c \ll 1$. The spectator picture would still apply at asymptotic quark masses, since $\Gamma ^{(1)}_{W-X}/ \Gamma _c\propto (f_D/\aver{E_{\bar q}})^2 \rightarrow 0$ as $m_c\rightarrow \infty$ due to $f_D \propto 1/\sqrt{m_c}$. Yet if eq.(3) were indeed to hold, it would have a dramatic impact on the theoretical description of weak heavy-flavour decays: the impact of this particular pre-asymptotic correction, namely WA, would be enhanced considerably and actually be quite significant even in beauty decays. Alternatively it had been suggested \cite{MINKOWSKI} that the wavefunction of the $D$ meson contains a $c \bar q g$ component where the $c \bar q$ pair forms a spin-one configuration with the gluon $g$ balancing the spin of the $c\bar q$ pair. \noindent Both effects, namely PI and WA, work in the same direction, i.e. both enhance $\tau (D^+)$ over $\tau (D^0)$. \noindent (iii) A rich structure emerges in the decays of charm baryons \cite{BARYONS1,BARYONS2,BARYONS3}: $\bullet$ On the one hand WS contributes to the Cabibbo allowed $\Lambda _c$ and $\Xi _c^0$ decays; one should also keep in mind that WS is {\em not} helicity suppressed in baryon decays already to lowest order in the strong coupling. It is still reduced in size by the corresponding wavefunction overlap; yet that is at least partially off-set by WS being described by two-body phasespace versus the three-body phase space of the spectator transition; this relative enhancement in phase space can be estimated to be roughly of order $16\pi ^2$. $\bullet$ PI affects the $\Lambda _c$, $\Xi _c^{0,+}$ and $\Omega _c$ widths in various ways, generating destructive as well as constructive contributions! This also strongly suggests that it is very hard to make reliable numerical predictions for these baryonic lifetimes; yet the overall qualitative pattern has been predicted: $$\tau (\Xi _c ^0) < \tau (\Lambda _c) < \tau (\Xi _c ^+) \eqno(4a)$$ together with $$\tau (\Lambda _c)< \tau (D^0) < \tau (D^+) \; . \eqno(4b)$$ \noindent (iv) Pre-asymptotic corrections might be sizeable in the lifetime ratios of beauty hadrons. \noindent Reviews of these phenomenological descriptions can be found in \cite{RUCKL,BRADLEE}. It turned out, as discussed in more detail later on, that some of the phenomenological descriptions anticipated the correct results: it is PI that provides the main engine behind the $D^+$-$D^0$ lifetime ratio; $\Lambda _c$ is considerably shorter-lived than $D^0$; the observed charm baryon lifetimes do obey the hierarchy stated in eq.(4). Nevertheless the phenomenological treatments had significant shortcomings, both of a theoretical and of a phenomenological nature: (i) No agreement had emerged in the literature about how corrections in particular due to WA and WS scale with the heavy quark mass $m_Q$. (ii) Accordingly no clear predictions could be made on the lifetime ratios among beauty hadrons, namely whether $\tau (B^+)$ and $\tau (B_d)$ differ by a few to several percent only, or by 20 - 30 \%, or by even more! (iii) No unequivocal prediction on $\tau (D_s)$ or $\tau (B_s)$ had appeared. (iv) In the absence of a systematic treatment it is easy to overlook relevant contributions, and that is actually what happened; or the absence of certain corrections had to be postulated in an ad-hoc fashion. Thus there existed an intellectual as well as practical need for a description based on a systematic theoretical framework rather than a set of phenomenological prescriptions. \subsection{From Phenomenology to Theory} In the last few years we have succeeded in showing that the non-perturbative corrections to heavy-flavour decays can be expressed through a {\em systematic} expansion in {\em inverse} powers of $m_Q$. A simple analogy with nuclear $\beta$ decay can illustrate this point. There are two effects distinguishing the decays of neutrons bound in a nucleus from the decay of free neutrons: \noindent (a) nuclear binding effects; \noindent (b) the impact of Pauli statistics correlating the electrons surrounding the nucleus with those emerging from $\beta$ decay. \noindent The typical energies of the bound electrons -- $\epsilon _{el}$ -- are certainly small compared to $E_{release}$, the energy released in the decay; let us assume -- although this is not true in reality -- that also the nuclear binding energies $\epsilon _{nucl}$ were small compared to $E_{release}$. In that case nuclear $\beta$ decays would proceed, to a good approximation, like the decays of {\em free} neutrons; corrections to this simple `spectator' picture could be computed via an expansion in powers of $\epsilon _{nucl}/E_{release}$ and $\epsilon _{el}/E_{release}$. In practice, however, the corrections for nuclear $\beta$ decay are incorporated by explicitely using the wave functions of the bound nucleons and electrons (obtained with the help of some fairly massive computer codes). There arise analogous corrections to the decay rate for a quark $Q$ inside a hadron $H_Q$: \noindent (a) interactions of the decaying quark with other partons in the hadron; this includes WA of $Q$ with the light valence antiquark for mesons or WS with the valence diquark system for baryons; they correspond to K capture of bound electrons by a heavy nucleus in the preceding example. \noindent (b) PI effects of the decay products with other partons in the hadron; e.g.: $b\bar u \rightarrow c\bar u d \bar u$ or $c\bar d \rightarrow u \bar d s \bar d$. They prolong the lifetimes of $D^+$ and $B^-$ mesons. \noindent The difference to the example of nuclear $\beta$ decay is quite obvious: even in the limit $m_Q\rightarrow \infty$ a non-relativistic bound-state treatment is inapplicable since the dynamical degrees of freedom of the heavy-flavour hadron $H_Q$ cannot fully be described by a hadronic wavefunction. The most reliable approach is then to evaluate weak decay rates of heavy-flavour hadrons through an expansion in powers of $\mu _{had}/m_Q$ where $\mu _{had}$ represents a hadronic scale $\leq$ 1 GeV. The first few terms in this series should yield a good approximation for beauty decays; the situation for charm decays is a priori unclear (and at present remains so a posteriori as well); this will be discussed in detail later on. \footnote{As already mentioned, top quarks decay weakly before they can hadronize \cite{RAPALLO}.} The vice of hadronization is then transformed into (almost) a virtue: the weak decays of heavy-flavour hadrons constitute an intriguing and novel laboratory for studying strong dynamics through their interplay with the weak forces -- and this is the secondary motivation for studying them. To be more specific: the heavy-flavour mass $m_Q$ provides an expansion parameter that allows to deal with the non-perturbative dynamics of QCD in a novel way. The formalism to be employed combines the $1/m_Q$ expansion with other elements derived from QCD proper without having to invoke a `deus ex machina' -- in contrast to phenomenological descriptions. There is one concept underlying, in one form or another, all efforts to deal with hadronization, namely the notion of quark-hadron duality (henceforth referred to as {\em duality} for short). In its broadest formulation it can be stated as follows: sufficiently inclusive transition rates between hadronic systems can be calculated in terms of quarks and gluons. While the general validity of this concept has not been established in a rigorous fashion, new light has been shed on its nature, validity and applicability by $1/m_Q$ expansions. Lifetimes obviously represent the most inclusive quantity where duality should be applicable. Nonleptonic transitions are certainly more complex than semileptonic ones; yet I will argue later that while there probably exists a quantitative difference in the degree to which duality holds in semileptonic and in nonleptonic decays, there is {\em no qualitative} one. Deatiled measurements of lifetimes are then theoretically important not only to translate data on the semileptonic branching ratio into a determination of the semileptonic width, but also in their own right. Dedicated and comprehensive studies of both charm and beauty decays are called for. The KM parameters relevant for charm decays -- $V(cs)$ and $V(cd)$ -- are well known through unitarity constraints of the 3x3 KM matrix, in contrast to the situation in beauty decays, and I consider it unlikely that charm decay studies can improve on that. On the other hand those can be used to calibrate our theoretical tools before applying them to beauty decays. Before concluding this general introduction I want to point out a less straightforward aspect of accurate lifetime measurements: decay rate evolutions in proper time for neutral mesons will not follow a single exponential function when particle-antiparticle oscillations occur. For there exist two distinct mass eigenstates with $\Delta m \equiv m_1-m_2\neq 0 \neq \Delta \Gamma \equiv \Gamma _1-\Gamma _2 $. The quantity $\Delta m$ generates a deviation of the form $e^{-\Gamma t}\cos \Delta mt$ or $e^{-\Gamma t}\sin \Delta mt$; $\Delta \Gamma \neq 0$ leads to the emergence of a second exponential. The general expression reads as follows: $$d\Gamma (B[D]\rightarrow f)/dt \; \propto e^{-\Gamma t}\cdot G(t)$$ $$G(t) = a+be^{-\Delta \Gamma t}+ ce^{-1/2\Delta \Gamma t}\cos \Delta mt + de^{-1/2\Delta \Gamma t}\sin \Delta mt\eqno(5a)$$ where $$a=|A(f)|^2\left[ \frac{1}{2} (1+|\frac{q}{p}\bar \rho (f)|^2) + Re[ \frac{q}{p}\bar \rho (f)] \right] $$ $$b=|A(f)|^2\left[ \frac{1}{2} [ 1+|\frac{q}{p}\bar \rho (f)|^2] - Re[ \frac{q}{p}\bar \rho (f)] \right] $$ $$c=|A(f)|^2\{ 1-|\frac{q}{p}\bar \rho (f)|^2\}\; , \; d= 2|A(f)|^2 Im[ \frac{q}{p}\bar \rho (f)] \; , \; \bar \rho (f)= \frac{\bar A(f)}{A(f)} \eqno(5b) $$ with $\bar A(f)$ and $A(f)$ denoting the amplitude for $\bar B [\bar D] \rightarrow f$ and $B [D] \rightarrow f$, respectively. \section{Preview of the Predictions on the Lifetime Ratios for Beauty and Charm Hadrons} In this section I summarize the numerical results and sketch the main elements of the underlying theoretical treatment in a way that can satisfy the casual reader. A more in-depth discussion of the theoretical concepts and tools will be presented in subsequent sections. Expanding the width for the decay of a heavy-flavour hadron $H_Q$ containing $Q$ into an inclusive final state $f$ through order $1/m_Q^3$ one obtains \cite{BUV,BS,SV} $$\Gamma (H_Q\rightarrow f)=\frac{G_F^2m_Q^5}{192\pi ^3}|KM|^2 \left[ c_3^f\matel{H_Q}{\bar QQ}{H_Q}_{norm}+ c_5^f\frac{ \matel{H_Q}{\bar Qi \sigma \cdot G Q}{H_Q}_{norm}}{m_Q^2}+ \right. $$ $$\left. +\sum _i c_{6,i}^f\frac{\matel{H_Q} {(\bar Q\Gamma _iq)(\bar q\Gamma _iQ)}{H_Q}_{norm}} {m_Q^3} + {\cal O}(1/m_Q^4)\right] , \eqno(6)$$ where the dimensionless coefficients $c_i^f$ depend on the parton level characteristics of $f$ (such as the ratios of the final-state quark masses to $m_Q$); $KM$ denotes the appropriate combination of KM parameters, and $\sigma \cdot G = \sigma _{\mu \nu}G_{\mu \nu}$ with $G_{\mu \nu}$ being the gluonic field strength tensor. The last term in eq.(6) implies also the summation over the four-fermion operators with different light flavours $q$. The expectation values of the local operators appearing on the right-hand side of eq.(6) contain the relativistic normalization of the state $|H_Q\rangle$: $$\matel{H_Q}{O_i}{H_Q}_{norm} \equiv \matel{H_Q}{O_i}{H_Q}/2M_{H_Q} \, .\eqno(7) $$ It is through the quantities $\matel{H_Q}{O_i}{H_Q}$ that the dependence on the {\em decaying hadron} $H_Q$, and on non-perturbative forces in general, enters, instead of through wavefunctions as in nuclear $\beta$ decay. Since these are matrix elements for on-shell hadrons $H_Q$, one sees that $\Gamma (H_Q\rightarrow f)$ is indeed expanded into a power series in $\mu _{had}/m_Q < 1$. For $m_Q\rightarrow \infty$ the contribution from the lowest dimensional operator obviously dominates; here it is the dimension three operator $\bar QQ$. A heavy quark expansion yields: $$\matel{H_Q}{\bar QQ}{H_Q}_{norm}=1+ {\cal O}(1/m_Q^2)\, , \eqno(8)$$ i.e. $\matel{H_Q}{\bar QQ}{H_Q}_{norm}=1$ for $m_Q\rightarrow \infty$, reflecting the unit of heavy-flavour common to all hadrons $H_Q$. Eqs.(6,8) show that the leading nonperturbative corrections are of order $1/m_Q^2$ rather than $1/m_Q$; therefore they can be expected to be rather small in beauty decays. The expectation values appearing in the $1/m_Q^2$ and $1/m_Q^3$ contributions can be related to other observables and their size thus be extracted in a reliable manner. Applying the master formula eq.(6) to lifetimes one arrives at the following general results: \noindent $\bullet$ The leading contribution to the total decay width is given by the first term on the right-hand-side of eq.(8) that is {\em common} to all hadrons of a given heavy-flavour quantum number. For $m_Q\rightarrow \infty$ one has thus derived -- from QCD proper -- the spectator picture attributing equal lifetimes to all hadrons of a given heavy-flavour! This is not a surprising result -- after all without hadronization there is of course only a unique lifetime --, but it is gratifying nevertheless. \noindent $\bullet$ Lifetime differences first arise at order $1/m_Q^2$ and are controlled by the expectation values of two dimension five operators, to be discussed later. These terms, which had been overlooked in the original phenomenological analyses, generate a lifetime difference between heavy-flavour {\em baryons} on one side and {\em mesons} on the other. Yet apart from small isospin or $SU(3)_{fl}$ breaking they shift the meson widths by the same amount and thus do {\em not} lead to differences among the meson lifetimes. \noindent $\bullet$ Those emerge at order $1/m_Q^3$ and are expressed through the expectation values of four-fermion operators which are proportional to $f_M^2$ with $f_M$ denoting the decay constant for the meson $M$. Contributions from what is referred to as WA and PI in the original phenomenological descriptions (see the discussion in Sect.1) are systematically and consistently included. Further contributions to the baryon-meson lifetime difference also arise at this level due to WS. \noindent $\bullet$ Since the transitions $b\rightarrow c l \nu$ or $c\rightarrow s l \nu$ are described by an isosinglet operator one can invoke the isospin invariance of the strong interactions to deduce for the semileptonic widths $$ \Gamma _{SL}(B^-)= \Gamma _{SL}(B_d) \eqno(9a)$$ $$ \Gamma _{SL}(D^+)= \Gamma _{SL}(D^0) \eqno(9b)$$ and therefore $$\frac{\tau (B^-)}{\tau (B_d)}= \frac{BR_{SL}(B^-)}{BR_{SL}(B_d)} \eqno(10a)$$ $$\frac{\tau (D^+)}{\tau (D^0)}= \frac{BR_{SL}(D^+)}{BR_{SL}(D^0)} \eqno(10b)$$ up to small corrections due to the KM [Cabibbo] suppressed transition $b\rightarrow u l \nu$ [$c\rightarrow d l \nu$] which changes isospin by half a unit. The spectator picture goes well beyond eqs.(9): it assigns the same semileptonic width to all hadrons of a given heavy flavour. Yet such a property cannot be deduced on {\em general} grounds: for one had to rely on $SU(3)_{Fl}$ symmetry to relate $\Gamma _{SL}(D_s)$ to $\Gamma _{SL}(D^0)$ or $\Gamma _{SL}(B_s)$ to $\Gamma _{SL}(B_d)$ and no symmetry can be invoked to relate the semileptonic widths of mesons and baryons. There is actually a WA process that generates semileptonic decays on the Cabibbo-allowed level for $D_s$ [and also for $B_c$], but not for $D^0$ and $D^+$ nor for $B_d$, $B^-$ and $B_s$ mesons: the hadrons are produced by gluon emission off the $\bar s$ [or the $\bar c$] line. Yet since the relative weight of WA is significantly reduced in meson decays, one does not expect this mechanism to change $\Gamma _{SL}(D_s)$ significantly relative to $\Gamma _{SL}(D^0)$. Actually there are corrections to the semileptonic widths arising already in order $1/m_Q^2$. On rather general grounds one predicts the expectation values $\matel{P_Q}{\bar QQ}{P_Q}$ and $\matel{P_Q}{\bar Qi\sigma \cdot G Q}{P_Q}$ to be largely independant of the flavour of the light antiquark in the meson and therefore $$\Gamma _{SL}(D_s)\simeq \Gamma _{SL}(D^0) \eqno(11a)$$ $$\Gamma _{SL}(B_s)\simeq \Gamma _{SL}(B_d) \eqno(11b)$$ like in the naive spectator picture, but for non-trivial reasons. On the other hand, as explained later, the values of the expectation values for these operators are different when taken between baryon states and one expects $$\Gamma _{SL}(\Lambda _Q) > \Gamma _{SL}(P_Q) \eqno(11c)$$ The remarks above can be summarized as follows: $$\Gamma (\Lambda _Q) = \Gamma ([Q\bar q]^0) = \Gamma ([Q\bar q]^{\pm}) + {\cal O}(1/m_Q^2) \eqno(12a)$$ $$\Gamma (\Lambda _Q) > \Gamma ([Q\bar q]^0) = \Gamma ([Q\bar q]^{\pm}) + {\cal O}(1/m_Q^3) \eqno(12b)$$ $$\Gamma (\Lambda _Q) > \Gamma ([Q\bar q]^0) > \Gamma ([Q\bar q]^{\pm}) + {\cal O}(1/m_Q^4) \eqno(12c)$$ Quantitative predictions for the lifetime ratios of beauty hadrons through order $1/m_b^3$ are given in Table \ref{TABLE1} together with present data. \begin{table} \begin{tabular} {|l|l|l|} \hline Observable &QCD ($1/m_b$ expansion) &Data \\ \hline \hline $\tau (B^-)/\tau (B_d)$ & $1+ 0.05(f_B/200\, \,\mbox{MeV} )^2 [1\pm {\cal O}(10\%)]>1$ &$1.04 \pm 0.05$ \\ &(mainly due to {\em destructive} interference) & \\ \hline $\bar \tau (B_s)/\tau (B_d)$ &$1\pm {\cal O}(0.01)$ & $ 0.98\pm 0.08$ \\ \hline $\tau (\Lambda _b)/\tau (B_d)$&$\sim 0.9\; ^*$ & $0.76\pm 0.06$ \\ \hline \end{tabular} \centering \caption{QCD Predictions for Beauty Lifetimes} \label{TABLE1} \end{table} The expectations \cite{MIRAGE,MARBELLA,DS} for the lifetimes of charm hadrons are juxtaposed to the data in Table \ref{TABLE2}. The overall agreement is rather good given the theoretical and experimental uncertainties. It has to be kept in mind that the {\em numerical} predictions on {\em baryon} lifetimes involve quark model results for the size of the relevant expectation values. This is indicated in Tables 1 and 2 by the asterisk. There is however one obvious discrepancy: the observed $\Lambda _b$ lifetimes is significantly shorter than predicted. While -- as indicated just above -- predictions on {\em baryon} lifetimes have to be taken with a grain of salt, one cannot change the prediction on $\tau (\Lambda _b)/\tau (B_d)$ with complete theoretical impunity; this will be explained later on. \begin{table} \begin{tabular} {|l|l|l|} \hline Observable &QCD ($1/m_c$ expansion) &Data \\ \hline \hline $\tau (D^+)/\tau (D^0)$ & $\sim 2 \; \; \; $ [for $f_D \simeq 200$ MeV] &$2.547 \pm 0.043$ \\ &(mainly due to {\em destructive} interference) & \\ \hline $\tau (D_s)/\tau (D^0)$ &$1\pm$ few $\times 0.01$ & $ 1.125\pm 0.042$ \\ \hline $\tau (\Lambda _c)/\tau (D^0)$&$\sim 0.5\; ^*$ & $0.51\pm 0.05$\\ \hline $\tau (\Xi ^+ _c)/\tau (\Lambda _c)$&$\sim 1.3\; ^*$ & $1.75\pm 0.36$\\ \hline $\tau (\Xi ^+ _c)/\tau (\Xi ^0 _c)$&$\sim 2.8\; ^*$ & $3.57\pm 0.91$\\ \hline $\tau (\Xi ^+ _c)/\tau (\Omega _c)$&$\sim 4\; ^* $& $3.9 \pm 1.7$\\ \hline \end{tabular} \centering \caption{QCD Predictions for Charm Lifetimes} \label{TABLE2} \end{table} These numerical predictions bear out the general expectations expressed above. This will be discussed below in more detail. For proper perspective on the role played by the heavy quark expansion some comments should be made at this point: \noindent (i) As stated before, the early phenomenological treatments had already identified the relevant mechanisms by which lifetime differences arise, namely PI, WA and WS. Furthermore the authors of ref.\cite{PI} had argued -- correctly -- that PI constitutes the dominant effect. Yet it is the heavy quark expansion that provides a firm theoretical underpinning to these expectations. In particular it clarifies -- both conceptually and quantitatively -- the role of WA and how its weight scales with $1/m_Q$. \noindent (ii) Contributions of order $1/m_Q$ would dominate all other effects -- if they were present! The heavy quark expansion shows unequivocally that they are absent in total rates; this has important ramifications, to be pointed out later. \noindent (iii) Corrections of order $1/m_Q^2$ differentiate between the decays of mesons and of baryons. They had been overlooked before. The heavy quark expansion has identified them and basically determined their size. \section{Methodology of the Heavy Quark Expansion for Fully Integrated Rates} The weak decay of the heavy quark $Q$ inside the heavy-flavour hadron $H_Q$ proceeds within a cloud of light degrees of freedom (quarks, antiquarks and gluons) with which $Q$ and its decay products can interact strongly. It is the challenge for theorists to treat these initial and final state hadronization effects. Among the existing four Post-Voodoo theoretical technologies -- QCD Sum Rules, Lattice QCD, Heavy Quark Effective Theory (HQET) and Heavy Quark Expansions -- only the last one deals with inclusive decays. On the other hand it benefits, as we will see, from the results of the other three technologies, since those can determine the size of some of the relevant expectation values.\footnote{It should be noted that contrary to frequent claims in the literature HQET -- as it is usually and properly defined -- per se does {\em not} allow to treat lifetimes or even the total semileptonic width: for those observables strongly depend on $m_Q$ whereas it is the special feature of HQET that $m_Q$ is removed from its Lagrangian.} In analogy to the treatment of $e^+e^-\rightarrow hadrons$ one describes the transition rate into an inclusive final state $f$ through the imaginary part of a forward scattering operator evaluated to second order in the weak interactions \cite{BUV,BS,SV}: $$\hat T(Q\rightarrow f\rightarrow Q)= i \, Im\, \int d^4x\{ {\cal L}_W(x){\cal L}_W^{\dagger}(0)\} _T\eqno(13)$$ where $\{ .\} _T$ denotes the time ordered product and ${\cal L}_W$ the relevant effective weak Lagrangian expressed on the parton level. If the energy released in the decay is sufficiently large one can express the {\em non-local} operator product in eq.(13) as an infinite sum of {\em local} operators $O_i$ of increasing dimension with coefficients $\tilde c_i$ containing higher and higher powers of $1/m_Q$.\footnote{It should be kept in mind, though, that it is primarily the {\em energy release} rather than $m_Q$ that controls the expansion.} The width for $H_Q\rightarrow f$ is then obtained by taking the expectation value of $\hat T$ between the state $H_Q$: $$\matel{H_Q}{\hat T (Q\rightarrow f\rightarrow Q)}{H_Q} \propto \Gamma (H_Q\rightarrow f) = G_F^2 |KM|^2 \sum _i \tilde c_i^{(f)} \matel{H_Q}{O_i}{H_Q} \eqno(14)$$ This master formula holds for a host of different inclusive heavy-flavour decays: semileptonic, nonleptonic and radiative transitions, KM favoured or suppressed etc. For semileptonic and nonleptonic decays treated through order $1/m_Q^3$ it takes the form already given in eq. (6): $$\Gamma (H_Q\rightarrow f)=\frac{G_F^2m_Q^5}{192\pi ^3}|KM|^2 \left[ c_3^f\matel{H_Q}{\bar QQ}{H_Q}_{norm}+ c_5^f\frac{ \matel{H_Q}{\bar Qi\sigma \cdot G Q}{H_Q}_{norm}}{m_Q^2}+ \right. $$ $$\left. +\sum _i c_{6,i}^f\frac{\matel{H_Q} {(\bar Q\Gamma _iq)(\bar q\Gamma _iQ)}{H_Q}_{norm}} {m_Q^3} + {\cal O}(1/m_Q^4)\right] $$ Integrating out the two internal loops in Figs.2 and 3 yields the operators $\bar QQ$ and $\bar Qi\sigma \cdot G Q$, respectively; the black boxes represent ${\cal L_W}$. Likewise the diagrams in Figs.4 generate four-quark operators; Fig.4a and 4b differ in how the light quark flavours are connected inside the hadron $H_Q$: cutting a quark line $q$ in Fig.2 and connecting it to the $\bar q$ constituent of the $H_Q$ meson, as shown in Fig.4a, one has a WA transition operator; cutting instead the $\bar q$ line and connecting it to the $H_Q$ constituents, see Fig.4b, leads to the four-fermion operator describing PI. Inspection of these diagrams suggests a rule-of-thumb that is borne out by explicit computations: the dimensionless coefficients $c^f_d$ are smaller for the two-loop diagrams of Figs.2 and 3 than for the one-loop diagrams of Figs.4. Integrating out the internal loops of Fig.3 actually generates not only the dimension-five chromomagnetic operator, but also dimension-six quark-gluon operators, namely $\bar Q (iD_{\mu}G_{\mu \nu})\gamma _{\nu}Q$ and $\bar Q\sigma _{\mu \nu}G_{\mu \rho}\gamma _{\rho} iD_{\nu}Q$. Using the equations of motion those can be reduced to four-quark operators. Yet their coefficients are suppressed by $\alpha _S/\pi$ relative to the coefficient coming from Figs.4; therefore one can discard their contributions at the present level of accuracy. One important general observation should be made at this point. The two operators $\bar QQ$ and $\bar Qi\sigma \cdot G Q$ obviously do not contain any light quark or antiquark fields; those enter first on the dimension-six level. Their expectation values are, as discussed below, practically insensitive to the light degrees of freedom up to terms of order $1/m_Q^3$. The inclusive semileptonic and non-leptonic widths through order $1/m_Q^2$ are thus controled by the {\em decay} of a single quark $Q$, yet occurring in a non-trivial background field; the {\em spectator} contribution universal to the widths for all hadrons $H_Q$ of a given heavy flavour constitutes its asymptotic part: $$\Gamma (H_Q) \simeq \Gamma _{decay}(H_Q) + {\cal O}(1/m_Q^3)\; , \eqno(15a)$$ $$\Gamma _{decay}(H_Q) \simeq \Gamma _{spect}(Q) + {\cal O}(1/m_Q^2)\; . \eqno(15b)$$ However $\Gamma _{decay}$ possesses a different value for mesons and baryons, as discussed below. The expansion introduced above will be of practical use only if its first few terms provide a good approximation and if one can determine the corresponding expectation values in a reliable manner. The latter can be achieved in two complementary ways: One relates the matrix element in question to other observables through a heavy quark expansion; or one harnesses other second-generation theoretical technologies, namely QCD sum rules, QCD simulations on the lattice and Heavy Quark Effective Theory, to calculate these quantities. The results obtained so far are listed below: (a) Using the equations of motion one finds for the leading operator $\bar QQ$: $$\bar QQ= \bar Q\gamma _0Q - \frac{\bar Q[(i\vec D)^2-(i/2)\sigma \cdot G]Q}{2m_Q^2}+ g_S^2\frac{\bar Q\gamma _0t^iQ \sum _q \bar q\gamma _0 t^iq}{4m_Q^3} + {\cal O}(1/m_Q^4)\eqno(16)$$ with the sum in the last term running over the light quarks $q$; the $t^i$ denote the colour $SU(3)$ generators. I have ignored total derivatives in this expansion since they do not contribute to the $H_Q$ expectation values. There emerge two dimension-five operators, namely $\bar Q (i\vec D)^2 Q$ and $\bar Q \sigma \cdot G Q$ with $\vec D$ denoting the covariant derivative. The first one represents the square of the spatial momentum of the heavy quark $Q$ moving in the soft gluon background and thus describes its kinetic energy\footnote{Since it is not a Lorentz scalar, it cannot appear in eq.(6).}. The second one constitutes the chromomagnetic operator that already appeared in eq.(6). The relevant dimension-six operators can be expressed as four-fermion operators. Since $\bar Q\gamma _0 Q$ constitutes the Noether current for the heavy-flavour quantum number one has $\matel{H_Q}{\bar Q\gamma _0Q}{H_Q}_{norm}=1$ leading to eq.(8). (b) The first non-perturbative corrections arise at order $1/m_Q^2$, and they can enter in two ways, namely through the chromomagnetic operator in the OPE of $\hat T$ or through the $1/m_Q$ expansion of the $H_Q$ expectation values of the local operator $\bar QQ$. (c) For the chromomagnetic operator one finds $$\matel{P_Q}{\bar Qi\sigma \cdot G Q}{P_Q}_{norm} \simeq \frac{3}{2} (M_{V_Q}^2-M_{P_Q}^2) \eqno(17a)$$ where $P_Q$ and $V_Q$ denote the pseudoscalar and vector mesons, respectively. For $\Lambda _Q$ and $\Xi _Q$ baryons one has instead $$\matel{\Lambda _Q}{\bar Qi\sigma \cdot G Q}{\Lambda _Q} \simeq 0 \simeq \matel{\Xi _Q}{\bar Qi\sigma \cdot G Q}{\Xi _Q}\eqno(17b)$$ since the light diquark system inside $\Lambda _Q$ and $\Xi _Q$ carries no spin. These baryons thus represent (though only through order $1/m_Q^2$) a simpler system than the mesons where the light antiquark of course carries spin. For $\Omega _Q$ baryons on the other hand the situation is different: for the light di-quarks $ss$ form a spin-one configuration. The $\Omega _Q$ expectation value is then given by the spin-splitting in the baryon masses: $$\matel{\Omega _Q}{\bar Qi\sigma \cdot G Q}{\Omega _Q}_{norm} \simeq \frac {4}{3} [M^2(\Omega _Q^{(3/2)}) - M^2(\Omega _Q)]\eqno(17c)$$ (d) We thus have $$\matel{\Lambda _Q}{\bar QQ}{\Lambda _Q}_{norm}=1 - \frac{\langle (\vec p_Q)^2\rangle _{\Lambda _Q}}{2m_Q^2}+ {\cal O}(1/m_Q^3)\eqno(18a)$$ $$\matel{P_Q}{\bar QQ}{P_Q}_{norm}=1 - \frac{\langle (\vec p_Q)^2\rangle _{P_Q}}{2m_Q^2}+ \frac{3}{8} \frac{M_{V_Q}^2-M_{P_Q}^2}{m_Q^2}+ {\cal O}(1/m_Q^3)\eqno(18b)$$ with the notation $\langle (\vec p_Q)^2\rangle _{H_Q} \equiv \matel {H_Q}{\bar Q (i\vec D)^2 Q}{H_Q}_{norm}$. The reason for the appearance of the kinetic energy term in eqs. (18) is quite transparent: The first two terms on the r.h.s. of eqs.(18) represent the mean value of the factor $\sqrt{1-\vec v^2}$ reflecting the time dilation slowing down the decay of the quark $Q$ moving inside $H_Q$. (e) The value of $\langle (\vec p_Q)^2\rangle _{H_Q}$ has not been determined yet in an accurate way (details will be given later); we know, though, it has to obey the inequality \cite{VOLOSHIN,OPTICAL} $$\langle (\vec p_Q)^2\rangle _{H_Q}\geq \frac{1}{2} \matel{H_Q}{\bar Qi\sigma \cdot G Q}{H_Q}_{norm} \equiv \aver{\mu _G^2}_{P_Q}\eqno(19)$$ which -- because of eq.(17) -- is useful only for $H_Q=P_Q$. \footnote{One should note that this inequality has now been derived in a fully field-theoretical treatment of QCD, rather than a merely quantum-mechanical one. Its proper interpretation will be discussed later.} On the other hand, the difference in the kinetic energy of Q inside baryons and mesons can be related to the masses of charm and beauty hadrons \cite{BUVPREPRINT}: $$ \langle (\vec p_Q)^2\rangle _{\Lambda _Q}- \langle (\vec p_Q)^2\rangle _{P_Q} \simeq \frac{2m_bm_c}{m_b-m_c}\cdot \{ [\langle M_B\rangle -M_{\Lambda _b}]- [\langle M_D\rangle -M_{\Lambda _c}] \} \eqno(20)$$ with $\langle M_{B,D}\rangle$ denoting the `spin averaged' meson masses: $$ \langle M_B\rangle \equiv \frac{1}{4}(M_B+3M_{B^*}) \eqno(21)$$ and likewise for $\langle M_D\rangle$. In deriving eq.(20) it was implicitely assumed that the $c$ quark can be treated also as heavy; in that case $\langle (\vec p_c)^2\rangle _{H_c} \simeq \langle (\vec p_b)^2\rangle _{H_b}$ holds. (f) Eqs.(17-18,20) show that the two {\em dimension-five} operators do produce differences in $P_Q$ vs. $\Lambda _Q/\Xi _Q$ vs. $\Omega _Q$ lifetimes (and semileptonic widths) of order $1/m_Q^2$; i.e., $\Gamma _{decay}(H_Q)$, as defined in eqs.(15), depends on whether $H_Q$ represents a (pseudoscalar) meson or a baryon. In the latter case it is sensitive to whether the light diquark system carries spin zero ($\Lambda _Q$, $\Xi _Q$) or one ($\Omega _Q$). The origin of such a difference is quite transparent. (i) The heavy quark $Q$ can be expected to possess a somewhat different kinetic energy inside a meson or a baryon; it can also be affected by spin-spin interactions with the light di-quarks. This difference in the heavy quark motion means that Lorentz time dilation prolongs the lifetime of the quark $Q$ to different degrees in baryons than in mesons. (ii) While there exists a spin interaction between $Q$ and the antiquark $\bar q$ in the meson or the $ss$ system in the $\Omega _Q$ baryon, there is no such coupling inside $\Lambda _Q$ or $\Xi _Q$. \noindent Yet these effects do not generate a significant difference among {\em meson} lifetimes since their expectation values satisfy isospin and $SU(3)_{fl}$ symmetry to a good degree of accuracy. (g) Differences in meson lifetimes are generated at order $1/m_Q^3$ by {\em dimension-six} four-quark operators describing PI as well as WA; the weight of the latter is, as explained in more detail in the next section, greatly reduced. The expectation values of these operators look very similar to the one controlling $B^0-\bar B^0$ oscillations: $$\matel{H_Q(p)}{(\bar Q_L\gamma _{\mu}q_L) (\bar q_L\gamma _{\nu}Q_L}{H_Q(p)}_{norm}\simeq \frac{1}{4m_{H_Q}}f^2_{H_Q}p_{\mu}p_{\nu}\eqno(22)$$ with $f_{H_Q}$ denoting the decay constant for the meson $H_Q$; the so-called bag factor has been set to unity, i.e. {\em factorization} has been assumed. (h) The situation becomes much more complex for $\Lambda _Q$ and baryon decays in general. To order $1/m_Q^3$ baryons lose the relative simplicity mentioned above: there are several different ways in which the valence quarks of the baryon can be contracted with the quark fields in the four-quark operators; furthermore WS is {\em not} helicity suppressed and thus can make a sizeable contribution to lifetime differences; also the PI effects can now be constructive as well as destructive. Finally one cannot take recourse to factorisation as a limiting case. Thus there emerge three types of numerically significant mechanisms at this order in baryon decays -- in contrast to meson decays where there is a {\em single} dominant source for lifetime differences -- and their strength cannot be expressed in terms of a single observable like $f_{H_Q}$. At present we do not know how to determine the relevant matrix elements in a model-independant way. The best available guidance and inspiration is to be gleaned from quark model calculations with their inherent uncertainties. This analysis had already been undertaken in the framework of phenomenological models \cite{BARYONS1,BARYONS2,BARYONS3} with the following qualitative results: \noindent -- WS contributes to $\Lambda _Q$ and $\Xi _Q^0$ decays; \noindent -- a {\em destructive} interference affects $\Lambda _Q$, $\Xi _Q$ and $\Omega _b$ transitions. \noindent -- a {\em constructive} interference enhances $\Xi _c$ and $\Omega _c$ decays. \noindent One thing should be obvious already at this point: with terms of different signs and somewhat uncertain size contributing to differences among baryon lifetimes one has to take even semi-quantitative predictions with a grain of salt! \noindent The probability amplitude for WS or interference to occur is expressed in quark models by the wavefunction for $Q$ and one of the light quarks at zero spatial separation. This quantity is then related to the meson wavefunction through the hyperfine splitting in the baryon and meson masses \cite{BARYONS3,MARBELLA}: $$\frac{|\psi _{\Lambda _Q}^{(Qq)}(0)|}{|\psi _{P_Q}(0)|}\simeq \frac{2m_q^*(M_{\Sigma _Q}-M_{\Lambda _Q})} {(M_{P_Q}-m_q^*)(M_{V_Q}-M_{P_Q})}\eqno(23)$$ with $m_q^*$ denoting the {\em phenomenological constituent} mass of the light quark $q$ to be employed in these models. For the baryonic expectation values one then obtains $$\matel{\Lambda _Q}{\bar Q\Gamma _{\mu}Q \bar q \Gamma _{\mu} q}{\Lambda _Q}_{norm} \sim \frac{1}{4\aver{\mu ^2_G}_{P_Q}}(M_{\Sigma _Q}-M_{\Lambda _Q}) m^*_q F^2_{P_Q}M_{P_Q}\eqno(24)$$ and likewise for $\Xi _Q$. For the $\Omega _Q$ one has instead $$\matel{\Omega _Q}{\bar Q\Gamma _{\mu}Q \bar q \Gamma _{\mu} q}{\Omega _Q}_{norm} \sim \frac{5}{6\aver{\mu ^2_G}_{P_Q}}(M_{\Sigma _Q}-M_{\Lambda _Q}) m^*_q F^2_{P_Q}M_{P_Q} \; . \eqno(25)$$ reflecting the different spin substructure which already had caused the difference in eqs.(17a) vs. (17b). For simplicity one has assumed in these expressions that the hyperfine splitting in the baryon masses are universal. To summarize this discussion: \noindent $\bullet$ The non-perturbative corrections to total widths through order $1/m_Q^3$ are expressed in terms of three non-trivial (types of) matrix elements: $\aver{\mu _G^2}_{H_Q}$, $\aver{(\vec p_Q)^2}_{H_Q}$ and $\matel{H_Q}{(\bar Q\Gamma _i q)(\bar q\Gamma _i Q)}{H_Q}$. \noindent $\bullet$ The size of $\aver{\mu _G^2}_{H_Q}$ is well-known for the mesons and $\Lambda _Q$ and $\Xi _Q$ baryons; reasonable estimates can be obtained for $\Omega _Q$ baryons. \noindent $\bullet$ A lower bound exists for the mesonic expectation value $\aver{(\vec p_Q)^2}_{P_Q}$, but not for the baryonic one. Various arguments strongly suggest that $\aver{(\vec p_Q)^2}_{H_Q}$ is close to its lower bound. Furthermore the quantity $\aver{(\vec p_Q)^2}_{\Lambda _Q}-\aver{(\vec p_Q)^2}_{P_Q}$, which is highly relevant for the difference in baryon vs. meson lifetimes (and semileptonic widths) can be extracted from mass measurements. \noindent $\bullet$ The expectation value of the four-quark operators between {\em mesons} can be expressed in terms of a single quantity, namely the decay constant $f_{P_Q}$. \noindent $\bullet$ However the {\em baryonic} expectation values of the four-quark operators constitute a veritable Pandora's box, which at present is beyond theoretical control. On the other hand measurements of inclusive weak decay rates and the resulting understanding of the role played by WS and PI present us with a novel probe of the internal structure of the heavy-flavour baryons. \section{Comments on the Underlying Concepts} Before explaining these predictions in more detail and commenting on the comparison with the data in the next subsection, I want to add four general remarks and elaborate on them. While I hope they will elucidate the underlying concepts, they are not truly essential for following the subsequent discussion and the more casual reader can ignore them. (A) {\em On the validity of the $1/m_Q$ expansion in the presence of gluon emission:} The WA contribution to the decay width of pseudoscalar mesons to lowest order in the strong coupling has been sketched in eq.(2). The diagram in Fig.1 contains gluon bremsstrahlung off the initial antiquark line in the $W$-exchange reaction; evaluating this diagram explicitely one finds, as stated before: $$\Gamma ^{(1)}_{W-X}\propto (\alpha_s /\pi ) G_F^2(f_{H_Q}/\aver{E_{\bar q}})^2m_Q^5 \eqno(26)$$ with $\aver {E_{\bar q}}$ denoting the average energy of the initial antiquark $\bar q$. Since $\Gamma ^{(1)}_{W-X} \propto 1/\aver{E_{\bar q}}^2$, it depends very sensitively on the low-energy quantity $ \aver{E_{\bar q}}$, and its magnitude is thus quite uncertain. The presence of a $1/\aver{E_{\bar q}}^2$ (or $1/m_q^2$) term would also mean that the nonperturbative corrections to even inclusive decay widths of heavy-flavour hadrons are not `infrared safe' and cannot be treated consistently through an expansion in inverse powers of the heavy quark mass $m_Q$ only. \noindent Fortunately the contribution stated in eq.(26) turns out to be spurious for {\em inclusive} transitions. This can best be seen by studying the imaginary part of the forward scattering amplitude $Q\bar q \rightarrow f \rightarrow Q\bar q$ as shown in Fig.5. There are actually three poles in this amplitude indicated by the broken lines which represent different final states: one with an on-shell gluon and the other two with an off-shell gluon materializing as a $q\bar q$ pair and involving interference with the spectator decay amplitude. These processes are all of the same order in the strong coupling and therefore have to be summed over for an {\em inclusive} decay. Analysing carefully the analyticity properties of the sum of these forward scattering amplitudes one finds \cite{MIRAGE,WA} that it remains finite in the limit of $\aver{E_{\bar q}} \rightarrow 0$, i.e. the amplitude for the {\em inclusive} width does {\em not} contain terms of order $1/\aver{E_{\bar q}}^2$ or even $1/\aver{E_{\bar q}}$! The contribution of WA can then be described in terms of the expectation value of a {\em local} four-fermion operator although the final state is dominated by low-mass hadronic systems. \noindent To summarize: the quantity $f_{H_Q}$ is calibrated by the large mass $m_Q$ rather than by $\aver{E_{\bar q}}$; WA is then only moderately significant even in charm decays. The $1/m_Q$ scaling thus persists to hold even in the presence of radiative corrections -- but only for {\em inclusive} transitions! This caveat is not of purely academic interest. For it provides a nice toy model illustrating the {\em qualitative} difference between inclusive and exclusive transitions. The latter are represented here by the three separate cuts in Fig.5. For charm decays they correspond to the reactions $[c\bar u] \rightarrow s \bar d g$ and $[c\bar u] \rightarrow s \bar d u\bar u$. The two individual transition rates quite sensitively depend on a low energy scale - $\aver{E_{\bar q}}$ - which considerably enhances the rate for the first transition while reducing it for the second one. Yet the dependance on $1/\aver{E_{\bar q}}$ disappears from their sum! (B) {\em On the fate of the corrections of order $1/m_Q$:} The most important element of eq.(2) is -- the one that is {\em missing}! Namely there is no term of order $1/m_Q$ in the total decay width whereas such a correction definitely exists for the mass formula: $M_{H_Q}=m_Q(1+\bar \Lambda /m_Q+{\cal O}(1/m_Q^2))$ (and likewise for differential decay distributions). Hadronization in the initial state does generate corrections of order $1/m_Q$ to the total width, as does hadronization in the final state. Yet local colour symmetry enforces that they cancel against each other! \footnote{This can be nicely illustrated in quantum mechanical toy models: The total rate for the transition $Q\rightarrow q$ depends on the {\em local} properties of the potential, i.e. on the potential around $Q$ in a neighourhood of size $1/m_Q$ only; the nature of the resulting spectrum for $q$ -- whether it is discrete or continuous, etc. -- depends of course on the long range properties of the potential, namely whether it is confining or not.} This can be understood in another more compact (though less intuitive) way as well: with the leading operator $\bar QQ$ carrying dimension three only dimension-four operators can generate $1/m_Q$ corrections; yet there is no independent dimension-four operator \cite{CHAY,BUV} {\em once the equation of motion is imposed} -- unless one abandons local colour symmetry thus making the operators $\bar Qi\gamma \cdot \partial Q$ and $\bar Qi\gamma \cdot B Q$ independent of each other ($B_{\mu}$ denotes the gluon field)! The leading non-perturbative corrections to fully integrated decay widths are then of order $1/m_Q^2$ and their size is controlled by two dimension-five operators, namely the chromomagnetic and the kinetic energy operators, as discussed above. Their contributions amount to no more than 10 percent for $B$ mesons -- $(\mu _{had}/m_b)^2$ $\simeq {\cal O}\left( (1\, GeV/m_b)^2\right) \sim {\cal O}(\% )$ -- as borne out by the detailed analysis previewed in Table 1. (C) {\em Which mass is it?} For a $1/m_Q$ expansion it is of course important to understand which kind of quark mass is to be employed there, in particular since for confined quarks there exists no a priori natural choice. It had been claimed that the pole mass can and therefore should conveniently be used. Yet such claims turn out to be fallacious \cite{POLEMASS,BRAUN}: QCD, like QED, is not Borel summable; in the high order terms of the perturbative series there arise instabilities which are customarily referred to as (infrared) renormalons representing poles in the Borel plane; they lead to an {\em additive} mass renormalization generating an {\em irreducible uncertainty} of order $\bar \Lambda$ in the size of the pole mass: $m_Q^{pole}=m_Q^{(0)}(1+c_1\alpha_s + c_2\alpha_s ^2 +...+ c_N\alpha_s ^N) + {\cal O}(\bar \Lambda )=m_Q^{(N)}(1+{\cal O}(\bar \Lambda/ m_Q^{(N)}))$. While this effect can safely be ignored in a purely perturbative treatment, it negates the inclusion of non-perturbative corrections $\sim {\cal O}(1/m_Q^2)$, since those are then parametrically smaller than the uncertainty $\sim {\cal O}(1/m_Q)$ in the definition of the pole mass. This problem can be taken care of through Wilson's prescriptions for the operator product expansion: $$\Gamma (H_Q\rightarrow f)=\sum _i c_i^{(f)}(\mu ) \matel{H_Q}{O_i}{H_Q}_{(\mu )}\eqno(27)$$ where a momentum scale $\mu$ has been introduced to allow a consistent separation of contributions from Long Distance and Short Distance dynamics -- $LD > \mu ^{-1} > SD$ -- with the latter contained in the coefficients $c_i^{(f)}$ and the former lumped into the matrix elements. The quantity $\mu$ obviously represents an auxiliary variable which drops out from the observable, in this case the decay width. In the limit $\mu \rightarrow 0$ infrared renormalons emerge in the coefficients; they cancel against ultraviolet renormalons in the matrix elements. Yet that does {\em not} mean that these infrared renormalons are irrelevant and that one can conveniently set $\mu =0$! For to incorporate both perturbative as well as non-perturbative corrections one has to steer a careful course between `Scylla' and `Charybdis': while one wants to pick $\mu \ll m_Q$ so as to make a heavy quark expansion applicable, one also has to choose $\mu _{had}\ll \mu$ s.t. $\alpha _S(\mu ) \ll 1$; for otherwise the {\em perturbative} corrections become uncontrollable. Wilson's OPE allows to incorporate both perturbative and non-perturbative corrections, and {\em this underlies also a consistent formulation of HQET}; the scale $\mu$ provides an infrared cut-off that automatically freezes out infrared renormalons. For the asymptotic difference between the hadron and the quark mass one then has to write $\bar \Lambda (\mu ) \equiv (M_{H_Q}-m_Q(\mu ))_{m_Q\rightarrow \infty}$. This nice feature does not come for free, of course: for one has to use a `running' mass $m_Q(\mu )$ evaluated at an intermediate scale $\mu$ which presents a technical complication. On the other hand this quantity can reliably be extracted from data \cite{OPTICAL}; furthermore it drops out from lifetime {\em ratios}, the main subject of this discussion. (D) {\em On Duality:} Quark-hadron duality equates transition rates calculated on the quark-gluon level with the observable ones involving the corresponding hadrons -- provided one sums over a sufficient number of final states. Since the early days of the quark model this concept has been invoked in many different formulations; among other things it has never been clearly defined what constitutes a sufficient number of final states. This lack of a precise formulation emerged since duality had never been derived from QCD in a rigorous fashion; furthermore a certain flexibility in an unproven, yet appealing intuitive concept can be of considerable heuristic benefit. \noindent Heavy quark expansions assume the validity of duality. Nevertheless they provide new and fruitful insights into its workings. As already stated, heavy quark expansions are based on an OPE, see eq.(26), and as such are properly defined in Euclidean space. There are two possible limitations to the procedure by which duality is implemented in heavy quark expansions. (i) The size of the coefficients $c_i^{(f)}$ is controlled by short-distance dynamics. In concrete applications they are actually obtained within perturbation theory. Those computations involve integrals over all momenta. However for momenta below the scale $\mu$ a perturbative treatment is unreliable. The contributions from this regime might turn out to be numerically insignificant for the problem at hand; yet in any case one can undertake to incorporate them through the expectation values of higher-dimensional operators -- the so-called condensates. This procedure is the basis of what is somestimes referred to as the `practical' version of the OPE. Yet it is conceivable that there are significant short-distance contributions from non-perturbative dynamics as well; instantons provide an illustration for such a complication, although their relevance is quite unclear at present \cite{MISHA}. (ii) Once the OPE has been constructed and its coefficients $c_i^f$ determined there, one employs a dispersion relation to analytically continue them into Minkowski space. This means, strictly speaking, that in Minskowski space only `smeared' transition rates can be predicted, i.e. transition rates averaged over some finite energy range. This situation was first analyzed in evaluating the cross section for $e^+e^- \rightarrow had $ \cite{POGGIO}. Through an OPE QCD allows to compute $$\aver{\sigma (e^+e^- \rightarrow had; E_{c.m.})} \equiv \frac{1}{\Delta E_{sm}} \int ^{E_{c.m.}+\Delta E_{sm}}_{E_{c.m.}-\Delta E_{sm}} dE'_{c.m.} \sigma (e^+e^- \rightarrow had; E'_{c.m.})$$ with $0< \Delta E_{sm} \ll E_{c.m.}$. If the cross section happens to be a smooth function of $E'_{c.m.}$ -- as it is the case well above production thresholds --, then one can effectively take the limit $\Delta E_{c.m.} \rightarrow 0$ to predict $\sigma (e^+e^- \rightarrow had; E_{c.m.})$ for a {\em fixed} c.m. energy $E_{c.m.}$. This scenario can be referred to as {\em local} duality. Yet close to a threshold, like for charm or beauty production, with its resonance structure one has to retain $\Delta E_{sm} \sim \mu _{had}\sim 0.5 - 1$ GeV. The same considerations are applied to heavy-flavour decays. Based on {\em global} duality one can predict `smeared' decay rates, i.e. decay rates averaged over a finite energy interval. If the energy release is large enough the decay rate will be a smooth function of it, smearing will no longer be required and {\em local} duality emerges. How large the energy release has to be for this simplification to occur cannot be predicted (yet). For the onset of local duality is determined by terms that cannot be seen in any finite order of the $1/m_Q$ expansion. However rather general considerations lead to the following expectations: (i) This onset should be quite abrupt around some energy scale exceeding usual hadronic scales. (ii) Local duality should hold to a good degree of accuracy for beauty decays, even for nonleptonic ones. (iii) On the other hand there is reason for concern that this onset might not have occurred for charm decays. This is sometimes referred to by saying that inclusive charm decays might receive significant contributions from (not-so-)`distant cuts'. \section{Size of the Matrix Elements} As discussed before, see eq.(27), the size of matrix elements depends in general on the scale $\mu$ at which they are evaluated. This will be addressed separately for the different cases. The scalar dimension-three operator $\bar QQ$ can be expanded in terms of $\bar Q\gamma _0Q$ -- with $\matel{H_Q}{\bar Q\gamma _0Q}{H_Q}_{norm} = 1$ -- and operators of dimension five and higher, see eq.(15). \noindent (i) Dimension-five operators \noindent Employing eq.(16) for $b$ and $c$ quarks one finds: $$\aver{\mu _G^2}_{B}\equiv \matel{B}{\bar b\frac{i}{2}\sigma \cdot Gb}{B}_{norm} \simeq \frac{3}{4} (M^2_{B^*}-M^2_B)\simeq 0.37\, (GeV)^2, \eqno(28a)$$ $$\aver{\mu _G^2}_D\equiv \matel{D}{\bar c\frac{i}{2}\sigma \cdot Gc}{D}_{norm} \simeq \frac{3}{4} (M^2_{D^*}-M^2_D)\simeq 0.41\, (GeV)^2,\eqno(28b)$$ i.e., these two matrix elements that have to coincide in the infinite mass limit are already very close to each other. They are already evaluated at the scale that is appropriate for beauty and charm decays. For the description of $B_c$ decays one needs the expectation value of the beauty as well as charm chromomagnetic operator taken between the $B_c$ state. While the $B_c^*$ and $B_c$ masses have not been measured yet, one expects the theoretical predictions \cite{QUIGG} on them to be fairly reliable. With $M(B^*_c)\simeq 6.33$ GeV and $M(B_c)\simeq 6.25$ GeV one obtains $$\matel{B_c}{\bar b\frac{i}{2}\sigma \cdot Gb}{B_c}_{norm} \simeq \matel{\bar B_c}{\bar c\frac{i}{2}\sigma \cdot Gc}{\bar B_c}_{norm} \simeq 0.75\; (GeV)^2 \; ; \eqno(29)$$ i.e., double the value as for the mesons with light antiquarks. {}From eq.(20) one obtains for the {\em difference} in the kinetic energy of the heavy quark inside baryons and mesons $$\langle (\vec p_b)^2\rangle _{\Lambda _b}- \langle (\vec p_b)^2\rangle _B \simeq -(0.07 \pm 0.20)\, (GeV)^2 \eqno(30)$$ using available data; for charm the same value holds to this order. We see that at present one cannot tell whether the kinetic energy of the heavy quark inside baryons exceeds that inside mesons or not. The uncertainty in the mass of $\Lambda _b$ constitutes the bottleneck; it would be quite desirable to decrease its uncertainty down to below 10 MeV or even less. For the overall size of $\langle (\vec p_Q)^2\rangle _{H_Q}$ we have the following lower bounds from eq.(19): $$\aver{(\vec p_b)^2}_B\geq 0.37 (GeV)^2\eqno(31)$$ $$\aver{(\vec p_c)^2}_D\geq 0.40 (GeV)^2\eqno(32)$$ To be conservative one can add a term of at most $\pm 0.15\; (GeV)^2$ to the right-hand side of eqs.(31,32) reflecting the uncertainty in the scale at which the expectation values are to be evaluated \cite{OPTICAL}. There are various arguments as to why the size of $\aver{(\vec p_b)^2}_B$ will not exceed this lower bound significantly. One should note that such values for $\aver{(\vec p_b)^2}_B$ are surprisingly large: for they state that the momenta of the heavy quark inside the hadron is typically around 600 MeV. An analysis based on QCD sum rules yields a value that is consistent with the preceding discussion \cite{QCDSR}: $$\aver{(\vec p_b)^2}_B = (0.5\pm 0.1)\; (GeV)^2\eqno(33)$$ It can be expected that practically useful values for $\aver{(\vec p_{c,b})^2}_{D,B}$ will be derived from lattice QCD in the forseeable future \cite{GUIDO}. The leading non-perturbative corrections are thus controlled by the following parameters: $$\frac{\langle \mu ^2_G\rangle _B}{m_b^2}\simeq 0.016 \; \; , \frac {\langle (\vec p_b)^2\rangle _B}{m_b^2} \sim 0.016 \eqno(34a)$$ $$\frac{\langle \mu ^2_G\rangle _D}{m_c^2}\simeq 0.21 \; \; , \frac {\langle (\vec p_c)^2\rangle _D}{m_c^2} \sim 0.21 \eqno(34b)$$ $$\frac{\aver{\mu ^2_G}_{\Omega _c}}{m_c^2} \simeq 0.12 \eqno(35)$$ The $\Omega _c$ mass hyperfine splitting has not been measured yet. A quark model estimate of 70 MeV has been used in evaluating eq.(17b). \noindent (ii) Dimension-six Operators \noindent The size of the decay constants $f_B$, $f_{B_s}$, $f_D$ and $f_{D_s}$ has not been determined yet with good accuracy. Various theoretical technologies yield $$f_D \sim 200\, \pm \, 30 \, MeV \eqno(36)$$ $$f_B \sim 180\, \pm \, 40 \, MeV \eqno(37)$$ $$\frac{f_{B_s}}{f_B} \simeq 1.1 - 1.2 \eqno(38)$$ $$\frac{f_{D_s}}{f_D} \simeq 1.1 - 1.2 \eqno(39)$$ where a higher reliability is attached to the predictions for the ratios of decay constants, i.e. eqs.(38,39), than for their absolute size, eqs.(36,37). Recent studies by WA75 and by CLEO on $D_s^+\rightarrow \mu ^+ \nu $ yielded values for $f_{D_s}$ that are somewhat larger, but still consistent with these predictions. It can be expected that lattice QCD will produce more precise results on these decay constants in the foreseeable future and that those will be checked by future measurements in a significant way. Yet there are two subtleties to be kept in mind here: \noindent ($\alpha$) The matrix element of the four-quark operator is related to the decay constant through the assumption of `vacuum saturation'. Such an ansatz cannot hold as an identity; it represents an approximation the validity of which has to depend on the scale at which the matrix element is evaluated. The quantities $f_B$ and $f_D$ are measured in $B\rightarrow \tau \nu ,\, \mu \nu$ and $D\rightarrow \tau \nu ,\, \mu \nu$, respectively and thus probed at the heavy-flavour mass, $m_B$ and $m_D$. Yet for the strong interactions controling the size of the expectation value of the four-quark operator the heavy-flavour mass is a completely foreign parameter. If vacuum saturation makes (approximate) sense anywhere, it has to be at ordinary hadronic scales $\mu \simeq 0.5 - 1$ GeV. That means the decay constant that is observed at the heavy-flavour scale has to be evaluated at the hadronic scale $\mu$; this is achieved by the so-called hybrid renormalization to be described later. \noindent ($\beta$) The difference in decay widths generated by the four-quark operators is stated as being proportional to $f^2_{H_Q}/m_Q^2$ with $f_{H_Q}= F_{H_Q}[1 - |\bar \mu |/m_Q +{\cal O}(1/m_Q^2)]$; $F_{H_Q}$ represents the leading or `static' term which behaves like $1/\sqrt{m_Q}$ for $m_Q\rightarrow \infty$ and therefore $F_D > F_B$. Thus asymptotically one has $F^2_{H_Q}/m_Q^2$ which indeed vanishes like $1/m_Q^3$. Yet the decay constant $f_{H_Q}$ -- partly as a consequence of its usual definition via an axialvector rather than a pseudoscalar current -- contains large pre-asymptotic corrections which lead to $f_D \sim f_B$. Then it is not clear which value to use for $f_{H_Q}$ when calculating width differences, the asymptotic one -- $F_{H_Q}$ -- or the one including the pre-asymptotic corrections -- $f_{H_Q}$ where numerically $F_{H_Q} > f_{H_Q}$ holds. . It can be argued that the main impact of some of the dimension-seven operators in the OPE for the meson decay width is to renormalise the $F^2_{H_Q}/m_Q^2$ term to $f^2_{H_Q}/m_Q^2$; yet this issue can be decided only through computing the contributions from dimension-seven operators. I will return to this issue when discussing beauty and charm decays specifically. Both these considerations also apply when determining the {\em baryonic} expectation values of four-quark operators \cite{MARBELLA}. Yet -- as already pointed out -- that is a considerably more murky affair, as will become quite apparent in the subsequent discussions of the beauty and charm lifetimes. While there are significant uncertainties and ambiguities in the values of the masses of beauty and charm quarks their difference which is free of renormalon contributions is tightly constrained: $$m_b-m_c\simeq \langle M\rangle _B - \langle M\rangle _D + \langle (\vec p)^2\rangle \cdot \left( \frac{1}{2m_c}- \frac{1}{2m_b}\right) \simeq 3.46 \pm 0.04\, GeV\, .\eqno(40)$$ This value agrees very well with the one extracted from an analysis of energy spectra in semileptonic $B$ decays \cite{VOLOSHIN2}. Lifetime ratios are hardly sensitive to this difference; we list it here mainly for completeness. \section{Predictions on Beauty Lifetimes} Within the Standard Model there is no real uncertainty about the weak forces driving beauty decays; at the scale $M_W$ they are given by the Lagrangian $${\cal L}_W^{\Delta B=1}(\mu = M_W)= \frac{4G_F}{\sqrt{2}} [V_{cb}\bar c_L\gamma _{\mu}b_L+ V_{ub}\bar u_L\gamma _{\mu}b_L]\cdot [V_{ud}^*\bar d_L\gamma _{\mu}u_L + V_{cs}^*\bar s_L\gamma _{\mu}c_L] \eqno(41)$$ for non-leptonic transitions with an analogous expression for semileptonic ones. In eq.(41) I have ignored Cabibbo-suppressed transitions and also the $b\rightarrow t$ coupling since here we are not interested in Penguin contributions. Radiative QCD corrections lead to a well-known renormalization at scale $m_b$, which is often referred to as {\em ultra-violet} renormalization: $${\cal L}_W^{\Delta B=1}(\mu = m_b)= \frac{4G_F}{\sqrt{2}} V_{cb}V_{ud}^* \{ c_1(\bar c_L\gamma _{\mu}b_L)(\bar d_L\gamma _{\mu}u_L) +c_2(\bar d_L\gamma _{\mu}b_L)(\bar c_L\gamma _{\mu}u_L) \} \eqno(42)$$ for $b\rightarrow c\bar ud$ transitions and an analogous expression for $b\rightarrow c \bar cs$; the QCD corrections are lumped together into the coefficients $c_1$ and $c_2$ with $$c_1=\frac{1}{2}(c_+ + c_-), \; \; c_2=\frac{1}{2}(c_+ - c_-) \eqno(43a)$$ $$c_{\pm}=\left[ \frac{\alpha _S(M_W^2)}{\alpha _S(m_b^2)} \right] ^{\gamma _{\pm}}, \; \; \gamma _+=\frac{6}{33-2N_f}=-\frac{1}{2}\gamma _- \eqno(43b)$$ in the leading-log approximation; $N_f$ denotes the number of active flavours. Numerically this amounts to: $$c_1(LL) \simeq 1.1, \; \; c_2(LL) \simeq -0.23 \, . \eqno(44a)$$ Next-to-leading-log corrections modify mainly $c_2$ \cite{CHERNYAK}: $$c_1(LL+NLL) \simeq 1.13, \; \; c_2(LL+NLL) \simeq -0.29 \, . \eqno(44b)$$ Radiative QCD corrections thus lead to a mild enhancement of the original coupling -- $c_1>1$ -- together with the appearance of an induced operator with a different colour flow: $c_2\neq 0$. Later on we will also include the so-called `hybrid' renormalization reflecting radiative corrections in the domain from $m_b$ down to $\mu _{had}$, the scale at which the hadronic expectation values are to be evaluated. Such perturbative corrections affect the overall scale of the semileptonic and non-leptonic width (and thus of the semileptonic branching ratio); but by themselves they cannot generate {\em differences} among the lifetimes of beauty hadrons. Those have to be initiated by non-perturbative contributions although their numerical size depends also on perturbative corrections. The semileptonic and non-leptonic widths of beauty {\em hadrons} $H_b$ through order $1/m_b^2$ are given by $$\frac{\Gamma _{SL,decay}(H_b)}{\Gamma _0} = \matel{H_b}{\bar bb}{H_b}_{norm}\cdot \left[ \eta _{SL}I_0(x,0,0) + \frac{\aver{\mu _G^2}_{H_b}}{m_b^2} (x\frac{d}{dx}-2)I_0(x,0,0)\right] \; , \eqno(45a)$$ $$\frac{\Gamma _{NL,decay}(H_b)}{\Gamma _0} = N_C\cdot \matel{H_b}{\bar bb}{H_b}_{norm}\cdot \left[ \frac{A_0}{3} [I_0(x,0,0)+ \rho _{c\bar c} I_0(x,x,0) +\right. $$ $$\left. \frac{\aver{\mu _G^2}_{H_b}}{m_b^2} (x\frac{d}{dx}-2)(I_0(x,0,0)+I_0(x,x,0))] - \frac{4}{3} A_2 \frac{\aver{\mu _G^2}_{H_b}}{m_b^2} \cdot [I_2(x,0,0,) + I_2(x,x,0)] \right] \; , \eqno(45b)$$ $$\Gamma _0 \equiv \frac{G_F^2m_b^5}{192 \pi ^3}|V(cb)|^2 \; . \eqno(45c)$$ The following notation has been used here: $I_0$ and $I_2$ are phase-space factors, namely $$I_0(x,0,0)= (1-x^2)(1-8x+x^2)-12x^2\log x \eqno(46a)$$ $$I_2(x,0,0)=(1-x)^3\; , \; \; x=(m_c/m_b)^2 \eqno(46b)$$ for $b \rightarrow c \bar ud/c l\bar \nu$ and $$I_0(x,x,0)=v(1-14x-2x^2-12x^3) + 24x^2(1-x^2) \log \frac{1+v}{1-v}\, , \, v=\sqrt{1-4x}\eqno(46c)$$ $$I_2(x,x,0)=v(1+\frac{x}{2}+3x^2)-3x(1-2x^2) \log \frac{1+v}{1-v} \, , \eqno(46d)$$ for $b\rightarrow c \bar cs$ transitions. The quantities $\eta _{SL}$, $\rho _{c\bar c}$, $A_0$ and $A_2$ represent the QCD radiative corrections. More specifically one has $$A_2=(c_+^2-c_-^2)\; \; \; , \; \; \; A_0= (c_-^2+2c_+^2)\cdot J \eqno(46e)$$ with $J$ reflecting the effect of the subleading logarithms \cite{PETRARCA} and $$\eta _{SL}\simeq 1+ \frac{2}{3}\frac{\alpha_s}{\pi}g(m_c/m_b,m_l/m_b,0) \; . \eqno(46f)$$ The function $g$ can be computed numerically for arbitrary arguments \cite{PHAM} and analytically for the most interesting case $m_l=0$ \cite{NIR}. The allowed values for $\rho _{c\bar c}$,which reflects the fact that QCD radiative corrections are quite sensitive to the final state quark masses, can be found in ref.\cite{BAGAN}. With $x\simeq 0.08$ one obtains $$I_0(x,0,0)|_{x=0.08}\simeq 0.56 \; \; , \; \; I_2(x,0,0)|_{x=0.08}\simeq 0.78 \; \; for \; \; b\rightarrow c \bar ud$$ $$I_0(x,x,0)|_{x=0.08}\simeq 0.24 \; \; , \; \; I_2(x,x,0)|_{x=0.08}\simeq 0.32 \; \; for \; \; b\rightarrow c \bar cs$$ Since these functions are normalized to unity for $x=0$, one notes that the final-state quark masses reduce the available phase space quite considerably. Some qualitative statements can illuminate the dynamical situation: \noindent (i) As indicated in eqs.(15), $\Gamma _{SL/NL,decay}$ differ from the naive spectator result in order $1/m_b^2$. \noindent (ii) Since $\, \bar bi\sigma \cdot G b\, $ and $\, \bar b (i\vec D)^2b\, $ are $SU(3)_{Fl}$ singlet operators and their expectation values are practically isospin and even $SU(3)_{Fl}$ invariant, one obtains, as stated before, $\tau (B_d)\simeq \tau (B^-)\simeq \bar \tau (B_s)$ through order $1/m_b^2$ and likewise $\tau (\Lambda _b)\simeq \tau (\Xi _b)$. Yet the meson and baryon lifetimes get differentiated on this level since $\aver{\mu _G^2}_B > 0 = \aver{\mu _G^2}_{\Lambda _b,\, \Xi _b}$ and $\aver{(\vec p_b)^2}_B \neq \aver{(\vec p_b)^2}_{\Lambda _b}$, see eqs.(20,30). \noindent (iii) The semileptonic branching ratios of $\Lambda _b$ and $\Xi _b$ baryons remain unaffected in order $1/m_b^2$ due to $\aver{\mu _G^2}_{\Lambda _b,\, \Xi _b}\simeq 0$, whereas for $B$ mesons it is (slightly) reduced \cite{BUV}; i.e., one finds $$BR_{SL}(B) < BR_{SL}(\Lambda _b) + {\cal O}(1/m_b^3)\; . $$ Beyond order $1/m_b^2$, however, this relation is changed, as discussed later on. \subsection{$B^-$ vs. $B_d$ Lifetimes} Differences in $\tau (B^-)$ vs. $\tau (B_d)$ are generated by the local dimension-six four-quark operators $(\bar b_L\gamma _{\mu}q_L) (\bar q_L\gamma _{\nu}b_L)$ which explicitely depend on the light quark flavours. Such corrections are of order $1/m_b^3$. Based on this scaling law one can already infer from the observed $D$ meson lifetime ratios that the various $B$ lifetimes will differ by no more than 10\% or so. Assuming factorization for the expectation values of these four-fermion operators, i.e. applying eq.(22), one obtains $$\frac{\Gamma (B_d)-\Gamma (B^-)}{\Gamma (B)} \sim \frac{\Gamma _{nonspect}(B)}{\Gamma _{spect}(B)} \propto \frac{f_B^2}{m_b^2}\; , \eqno(47)$$ as discussed above. It turns out that WA can change the $B$ lifetimes by no more than1\% or so; due to interference with the spectator reaction, they could conceivably {\em prolong} the $B_d$ lifetime relative to the $B^-$ lifetime, rather than reduce it. The dominant effect is clearly provided by PI, which produces a negative contribution to the $B^-$ width: $$\Gamma (B^-)\simeq \Gamma _{spect}(B)+ \Delta \Gamma _{PI}(B^-)\eqno(48a)$$ $$ \Delta \Gamma _{PI}(B^-)\simeq \Gamma _0\cdot 24\pi ^2\frac{f_B^2}{m_b^2} \left[ c_+^2-c_-^2+\frac{1}{N_C}(c_+^2+c_-^2)\right] \; .\eqno(48b)$$ Eq.(48b) exhibits an intriguing result: $\Delta \Gamma _{PI}(B^-)$ is positive for $c_+=1=c_-$, i.e. PI acts {\em constructively} and thus would shorten the $B^-$ lifetime in the {\em absence} of radiative QCD corrections \footnote{This shows that the term of `Pauli Interference' should {\em not} be construed as implying that the interference is a priori destructive.}. Including those as evaluated on the leading-log and next-to-leading-log level -- $c_+\simeq 0.84$, $c_-\simeq 1.42$ -- turns PI into a {\em destructive} interference prolonging the $B^-$ lifetime, albeit by a tiny amount only at this point. In eq.(48b) only ultraviolet renormalization has been incorporated. Hybrid renormalization \cite{HYBRID} down to the hadronic scale $\mu _{had}$ amplifies this effect considerably; one obtains $$ \Delta \Gamma _{PI}(B^-)\simeq \Gamma _0\cdot 24\pi ^2\frac{f_B^2}{M_B^2}\kappa ^{-4} \left[ (c_+^2-c_-^2)\kappa ^{9/4}+\frac{c_+^2+c_-^2}{3} -\frac{1}{9}(\kappa ^{9/2}-1)(c_+^2-c_-^2)\right] \, , $$ $$\kappa \equiv \left[ \frac{\alpha_s (\mu _{had}^2)} {\alpha_s (m_b^2)}\right] ^{1/b}, \; b=11-\frac{2}{3} n_F \eqno(49)$$ Putting everything together one finds: $$\frac{\tau (B^-)}{\tau (B_d)}\simeq 1+0.05 \cdot \frac{f_B^2}{(200\; MeV)^2}\; , \eqno(50)$$ i.e. the lifetime of a charged $B$ meson is predicted to definitely {\em exceed} that of a neutral $B$ meson by typically several percent. Three comments are in order for properly evaluating eq.(49): \noindent $\bullet$ Although WA could conceivably prolong the $B_d$ lifetime as stated before, its numerical significance pales by comparison to that of PI. PI acts destructive in $B^-$ decays once radiative QCD corrections are included to the best of our knowledge. The lifetime of $B^-$ mesons therefore has to exceed that of $B_d$ mesons. \noindent $\bullet$ While the sign of the effect is predicted in an unequivocal manner, its magnitude is not. The main uncertainty in the prediction on $\tau (B^-)/\tau (B_d)$ is given by our present ignorence concerning the size of $f_B$. \noindent $\bullet$ Even a precise measurement of the lifetime ratio $\tau (B^-)/\tau (B_d)$ would not automatically result in an exact determination of $f_B$ by applying eq.(50). For corrections of order $1/m_b^4$ have not (or only partially -- see the discussion in the preceding subsection) been included. Unless those have been determined, one cannot extract the size of $f_B$ even from an ideal measurement with better than a roughly 15 \% uncertainty. \subsection{$B_s$ Lifetimes} Very little $SU(3)_{Fl}$ breaking is anticipated between the $B_d$ and $B_s$ expectation values of the two dimension-five operators. \footnote{This issue will be addressed in some detail in the later discussion of $D^0$ vs. $D_s$ lifetimes.} Among the contributions from dimension-six operators WA affects $B_d$ and $B_s$ lifetimes somewhat differently due to different colour factors for these two decays. Yet it is quite irrelevant in either case. Thus one predicts the $B_d$ and the average $B_s$ lifetimes to practically coincide: $$\tau (B_d) \simeq \bar \tau (B_s) \pm {\cal O}(1\%) \eqno(51)$$ The term `average $B_s$ lifetime' is used for a reason: $B_s - \bar B_s$ oscillations generate two neutral beauty mesons carrying strangeness that differ in their mass as well as in their lifetime. Due basically to $m_t^2 \gg m_c^2$ one finds $\Delta m(B_s) \gg \Delta \Gamma (B_s) \neq 0$. While $\Delta m(B_s)$ can be calculated in terms of the expectation value of the {\em local} four-fermion operator $(\bar b_L\gamma _{\mu }q_L)(\bar b_L\gamma _{\mu }q_L)$ (since $m_b \ll m_t$), the situation is more complex for $\Delta \Gamma (B_s)$, since the underlying operator is nonlocal. One can however apply a heavy quark expansion; to lowest nontrivial order one obtains \cite{BSBS} $$\frac{\Delta \Gamma (B_s)}{\bar \Gamma (B_s)} \equiv \frac{\Gamma (B_{s,short})-\Gamma (B_{s,long})} {\bar \Gamma (B_s)}\simeq 0.18\cdot \frac{(f_{B_s})^2}{(200 \, MeV)^2} \eqno(52)$$ for $f_{B_s}$ not much larger than 200 MeV. Comparing eqs.(50-52) leads to the intriguing observation that the largest lifetime difference among $B^-$, $B_d$ and $B_s$ mesons is generated by a very subtle source: $B_s - \bar B_s$ oscillations! There are also two different lifetimes for neutral $B$ mesons without strangeness; yet $\Delta \Gamma (B_d)$ is suppressed by $\sim \sin \theta ^2_C$ relative to $\Delta \Gamma (B_s)$. One can search for the existence of two different $B_s$ lifetimes by comparing $\tau (B_s)$ as measured in $B_s \rightarrow \psi \eta /\psi \phi$ on one hand and in $B_s\rightarrow l \nu X$ on the other. The former decay predominantly leads to a CP even final state and thus would, to good accuracy, reveal $\tau (B_{s,short})$; the latter exhibits the average lifetime $\bar \tau (B_s)= [\tau (B_{s,long} - \tau (B_{s,long})]/2$; for semileptonic $B_s$ decays involve the CP even and odd components in a nearly equal mixture. Thus $$\tau (B_s \rightarrow l \nu D^{(*)}) - \tau (B_s\rightarrow \psi \eta /\psi \phi ) \simeq \frac{1}{2} [\tau (B_{s,long})-\tau (B_{s,short})]\simeq $$ $$\simeq 0.09\cdot \frac{(f_{B_s})^2}{(200 \, MeV)^2}. \eqno(53)$$ Whether an effect of this size is large enough to be ever observed in a real experiment, is doubtful. Nevertheless one should search for it even if one has sensitivity only for a 50\% lifetime difference or so. For while eq.(52) represents the best presently available estimate, it is not a `gold-plated' prediction. It is conceivable that the underlying computation {\em underestimates} the actual lifetime difference! \subsection{$B_c$ Lifetime} $B_c$ mesons with their two heavy constituents -- $B_c=(b\bar c)$ -- represent a highly intriguing system that merits special efforts to observe it. One expects a rich spectroscopy probing the inter-quark potential at distances intermediate to those that determine quarkonia spectroscopy in the charm and in the beauty system. The Isgur-Wise function for the striking channel $B_c\rightarrow l \nu \psi$ can be computed. What is most relevant for our discussion here is that its overall decays, and thus its lifetime, reflect the interplay of three classes of transitions, namely the decay of the $b$ quark, that of the $\bar c$ (anti)quark and WA of $b$ and $\bar c$: $$\Gamma (B_c)\simeq \Gamma _{b\rightarrow c} (B_c) + \Gamma _{\bar c \rightarrow \bar s}(B_c) + \Gamma _{WA}(B_c) \eqno(54)$$ While $b\rightarrow c$ and $\bar c \rightarrow \bar s$ transitions do not interfere with each other in any practical way and one can thus cleanly separate their widths, the situation is much more delicate concerning $\Gamma _{b \rightarrow c}$ and $\Gamma _{WA}$, as briefly explained later. Yet for the moment I ignore the latter although its helicity and wavefunction suppressions represented by the factors $m_c^2/m_b^2$ and $f^2_{B_c}/m_b^2$ are relatively mild ($f_{B_c} \simeq 400-600$ MeV \cite{QUIGG}) and partially offset by the numerical factor $16\pi ^2$ reflecting the enhancement of two-body phase space -- relevant for WA -- over three-body phase space appropriate for the spectator decay. Concerning the other two transitions one would naively expect the $\bar c$ decay to dominate: $\Gamma _{\bar c\rightarrow \bar s}(B_c) \sim \Gamma (D^0) \simeq (4\times 10^{-13}\, sec)^{-1} > \Gamma _{b \rightarrow c}(B_c) \sim \Gamma (B) \simeq (1.5\times 10^{-12}\, sec)^{-1}$. It had been suggested that for a tightly bound system like $B_c$ the decay width should be expressed not in terms of the usual quark masses (whatever they are), but instead in terms of effective quark masses reduced by something like a binding energy $\mu _{BE} \sim 500$ MeV. {\em If} so, then the $B_c$ width would be reduced considerably since due to $\Gamma _Q \propto m_Q^5$ one would find a relative reduction $(\Gamma _Q + \Delta \Gamma _Q)/\Gamma _Q \sim (m_Q - \mu _{BE})^5/m_Q^5 \sim 1 -5\mu _{BE}/m_Q$. Even more significantly, beauty decays would become more abundant than charm decays in the $B_c$ transitions since the binding energy constitutes a higher fraction of $m_c$ than of $m_b$: $\mu _{BE}/m_c > \mu _{BE}/m_b$. However this conjecture that might look quite plausible at first sight, turns out to be fallacious! For it is manifestly based on the existence of nonperturbative corrections of order $1/m_Q$; yet as discussed in Sect.4 those are absent in fully integrated decay rates like lifetimes due to a non-trivial cancellation between $1/m_Q$ contributions from initial-state and final-state radiation! The leading corrections arise in order $1/m_Q^2$ and they enter through the first two terms on the right-hand-side of eq.(54): $$\Gamma _{b \rightarrow c}(B_c) = \Gamma _{b \rightarrow c,decay}(B_c) + {\cal O}(1/m_b^3) \eqno(55a)$$ $$\Gamma _{\bar c \rightarrow \bar s}(B_c) = \Gamma _{\bar c \rightarrow \bar s,decay}(B_c) + {\cal O}(1/m_c^3) \eqno(55b)$$ The quantities $\Gamma _{b \rightarrow c,decay}(B_c)$ and $\Gamma _{\bar c \rightarrow \bar s}(B_c)$ are defined as in eqs.(45) and (68); i.e., the differences between $\Gamma _{decay}(B)$ and $\Gamma _{b \rightarrow c,decay}(B_c)$ enter through the different expectation values of the same operators $\bar QQ$ and $\bar Q\frac{i}{2}\sigma \cdot G Q$, and likewise for $\Gamma _{decay}(D)$ (to be discussed in detail in Sect.7) vs. $\Gamma _{\bar c \rightarrow \bar s,decay}(B_c)$. For the $B_c$ meson provides a different environment for quark decay than either $B$ or $D$ mesons. The relevance of that difference is illustrated by eqs.(28,29). Similarly one expects $\aver{(\vec p_b)^2}_{B_c}$, $\aver{(\vec p_c)^2}_{B_c}$ to differ from $\aver{(\vec p_b)^2}_B$, $\aver{(\vec p_c)^2}_D$. While there are large corrections of order $1/m_c^2$ in $\Gamma _{\bar c\rightarrow \bar s,decay}(B_c)$ that reduce the corresponding semileptonic branching ratio considerably, they largely cancel against each other in the total width. Therefore $$ \Gamma _{\bar c\rightarrow \bar s,decay}(B_c) \sim \Gamma _{decay}(D) \simeq \Gamma (D^0) \eqno(56)$$ Writing the $\Delta B=1$ width of the $B_c$ meson as a simple incoherent sum $\Gamma _{b\rightarrow c}+\Gamma _{WA}$ actually represents an oversimplification. For there arises considerable interference between higher order WA and spectator processes. Yet for the purposes of our discussion here such effects can be ignored; they will be discussed in detail in a forthcoming publication. The $1/m_b^2$ contributions to $\Gamma _{b\rightarrow c,decay}(B_c)$ are small, namely around a few percent. The numerical impact of $\Gamma _{WA}(B_c \rightarrow X_{\bar c s})$, which is formally of order $1/m_b^3$, is nevertheless sizeable due to the large decay constant, the merely mild helicity suppression, given by $m_c^2/m_b^2$, and the enhancement of WA by a factor $\sim 16\pi ^2$ due to its two-body kinematics. One finds: $$\Gamma _{b \rightarrow c}(B_c) \geq \Gamma _{b\rightarrow c}(B)\; . \eqno(57)$$ Comparing eqs.(56,57) one concludes that the naive expectation turns out to be basically correct, i.e. $\tau (B_c)$ is well below $10^{-12}$ sec and charm decays dominate over beauty decays with an ensuing reduction in the `interesting' branching ratios like $B_c\rightarrow l\nu \psi$ or $B_c\rightarrow \psi \pi$. \subsection{Beauty Baryon Lifetimes} It is quite natural to assume that the kinetic energy of the $b$ quark is practically the same inside the $\Lambda _b$ and $\Xi _b$ baryons: $$\aver{(\vec p _b)^2}_{\Lambda _b}\simeq \aver{(\vec p _b)^2}_{\Xi _b} \eqno(58a)$$ Together with eqs.(15,17a) this yields $$\matel{\Lambda _b}{\bar bb}{\Lambda _b}= \matel{\Xi _b}{\bar bb}{\Xi _b} + {\cal O}(1/m_b^3) \eqno(58b)$$ The leading {\em differences} among the $\Lambda _b$ and $\Xi _b$ lifetimes then arise in order $1/m_b^3$: they are generated by four-quark operators analogous to those that had already been identified in the phenomenological studies of charm baryons\cite{BARYONS1,BARYONS2,BARYONS3}. Some complexities arise, though, due to the presence of the two transitions $b\rightarrow c \bar ud$ and $b\rightarrow c \bar cs$; one finds $$\Gamma (\Lambda _b) = \Gamma _{decay}(\Lambda _b) + \Gamma _{WS}(\Lambda _b) -|\Delta \Gamma _{PI,-}(\Lambda _b, b\rightarrow c\bar ud)| \eqno(59a)$$ $$\Gamma (\Xi ^0 _b) = \Gamma _{decay}(\Xi _b) + \Gamma _{WS}(\Xi _b) - |\Delta \Gamma _{PI,-}(\Xi _b, b\rightarrow c\bar cs)| \eqno(59b)$$ $$\Gamma (\Xi ^- _b) = \Gamma _{decay}(\Xi _b) -|\Delta \Gamma _{PI,-}(\Xi _b, b\rightarrow c \bar ud)| -|\Delta \Gamma _{PI,-}(\Xi _b, b\rightarrow c\bar cs)| \eqno(59c)$$ where \footnote{The channel $b \rightarrow \tau \nu q$ has been ignored here for simplicity.} $$\Gamma _{decay}(\Lambda _b/\Xi _b) \equiv 2 \Gamma _{SL,decay}(\Lambda _b/\Xi _b)+ \Gamma _{NL,decay}(\Lambda _b/\Xi _b) \eqno(60a)$$ $$\frac{\Gamma _{SL,decay}(\Lambda _b/\Xi _b)}{\Gamma _0} = \matel{\Lambda _b/\Xi _b}{\bar bb}{\Lambda _b/\Xi _b}_{norm}\cdot \eta _{SL}I_0(x,0,0) \eqno(60b)$$ $$\frac{\Gamma _{NL,decay}(\Lambda _b/\Xi _b)}{\Gamma _0} = \matel{\Lambda _b/\Xi _b}{\bar bb}{\Lambda _b/\Xi _b}_{norm}\cdot A_0\cdot [I_0(x,0,0)+ \rho _{c\bar c} I_0(x,x,0)] \eqno(60c)$$ using the same notation as in eqs.(45); $$\Gamma _{WS}(\Lambda _b, \Xi _b) \equiv 2\tilde \Gamma _0 c_-^2\matel{\Lambda _b/\Xi _b}{\bar b_L\gamma _{\mu}b_L \bar u_L \gamma _{\mu}u_L}{\Lambda _b/\Xi _b}_{norm} \eqno(60d)$$ $$\frac{\Delta \Gamma _{PI,-}(\Lambda _b, b\rightarrow c\bar ud)} {\tilde \Gamma _0} \equiv - c_+(2c_- - c_+) \matel{\Lambda _b} {\bar b_L\gamma _{\mu}b\bar d_L\gamma _{\mu}d_L + \frac{2}{3}\bar b\gamma _{\mu}\gamma _5b \bar d_L\gamma _{\mu}d_L} {\Lambda _b}_{norm} \eqno(60e)$$ $$\frac{\Delta \Gamma _{PI,-}(\Xi _b, b\rightarrow c\bar cs)} {\tilde \Gamma _0}\equiv - c_+(2c_- - c_+) \matel{\Xi _b} {\bar b_L\gamma _{\mu}b\bar s_L\gamma _{\mu}s_L + \frac{2}{3}\bar b\gamma _{\mu}\gamma _5b \bar s_L\gamma _{\mu}s_L} {\Xi _b}_{norm} \eqno(60f)$$ $$\tilde \Gamma _0 \equiv 48\pi ^2 \Gamma _0 \eqno(60g)$$ $\Delta \Gamma _{PI,-}(\Xi _b, b\rightarrow c\bar ud)$ is obtained from eq.(60e) by taking the expectation value between $\Xi _b$ rather than $\Lambda _b$ states. Those expectation values vanish for the wrong charge-flavour combination; i.e., $\matel{\Xi _b^0}{\bar b_L\gamma _{\mu}b\bar d_L\gamma _{\mu}d_L} {\Xi _b^0}=0= \matel{\Lambda _b}{\bar b_L\gamma _{\mu}b\bar s_L\gamma _{\mu}s_L} {\Lambda _b}$. $\Gamma _{decay} $ includes the naive spectator term: $\Gamma _{decay} = \Gamma _{spect} + 1/m_b^2$ contributions; the latter are practically identical for $\Lambda _b$ and $\Xi _b$, but differ for $B$ mesons. $\Gamma _{WS}$ denotes the contribution due to WS and $\Delta \Gamma _{PI,-}$ the reduction due to destructive interference in the channels $b\rightarrow c\bar ud$ and $b\rightarrow c \bar cs$, respectively. \footnote{When $|\Delta \Gamma _{PI,-}(b\rightarrow c \bar ud)|$ is used in eq.(59c), one has, strictly speaking, to evaluate the expectation value for the state $\Xi _b$.} On general grounds one thus obtains an inequality: $$\tau (\Xi _b^-) > \tau (\Lambda _b)\, , \, \tau (\Xi _b^0) \eqno(61)$$ Naively one might expect $\Gamma (\Xi _b^0) > \Gamma (\Lambda _b)$ to hold, since the $b\rightarrow c \bar cs$ part of the width which is reduced by PI in $\Xi _b^0$ decays is smaller than the $b\rightarrow c \bar ud$ component which suffers from PI reduction in $\Lambda _b$ decays. On the other hand since phase space is more restricted for $b\rightarrow c \bar cs$ than for $b\rightarrow c \bar ud$, one would likewise expect the degree of (destructive) interference to be higher for the former than the latter; it is then quite conceivable that actually $\Gamma (\Xi _b^0) < \Gamma (\Lambda _b)$ holds. The {\em sign} of the difference in the $\Lambda _b$ and $\Xi _b^0$ lifetimes therefore provides us with valuable information on the strong dynamics. There are two complementary ways to transform these qualitative predictions into quantitative ones. \noindent (i) One evaluates the required expectation values explicitely within a quark model, as expressed in eq.(17b, 23-25). Since the model also predicts the baryon masses in terms of its parameters, one can cross check it with the observed spectroscopy. This will become increasingly relevant in the future, yet at present provides little guidance. The expressions given in eqs.(23-25) have to be augmented by the radiative QCD corrections: $$\matel{\Lambda _b}{\bar b\Gamma _{\mu}b \bar q \Gamma _{\mu} q}{\Lambda _b}_{norm} \sim \frac{1}{4\aver{\mu ^2_G}_B}(M_{\Sigma _b}- M_{\Lambda _b}) m^*_q F^2_BM_B\cdot \kappa ^{-4} \eqno(62a)$$ $$\matel{\Omega _b}{\bar b\Gamma _{\mu}b \bar q \Gamma _{\mu} q}{\Omega _b}_{norm} \sim \frac{1}{4\aver{\mu ^2_G}_B}(M_{\Sigma _b}- M_{\Omega _b}) m^*_q F^2_BM_B\cdot \kappa ^{-4} \eqno(62b)$$ with $\kappa$ as defined in eq.(49). It should be noted that here -- unlike in the case of meson decays -- the sign of the PI contribution is quite robust under radiative corrections: it is proportional to $- c_+(2c_- - c_+)$ which is negative already for $c_+=1=c_-$. Furthermore $\Gamma _{WS}$ is proportional to $c_-^2$ and thus colour-enhanced, since the baryon wavefunction is purely antisymmetric in colour space. This prescription yields lifetime differences of not more than a few percent in eq.(59) and $$ \frac{\tau (\Lambda _b)}{\tau (B_d)} \simeq 0.9 - 0.95 \; . \eqno(63a)$$ (ii) As will be discussed in the next subsection, the pattern predicted for charm baryon lifetimes agrees with the observations within the experimental and theoretical uncertainties. Taking this as prima facie evidence that the heavy quark expansion through order $1/m_Q^3$ applies -- at least in a semi-quantitative fashion -- already to charm lifetime ratios, one can extrapolate to the weight of these pre-asymptotic corrections in beauty decays using scaling like $1/m_Q^2$ and $1/m_Q^3$ (or $f_M^2/m_Q^2$). That way one again finds that the differences in eq.(56) amount to not more than a few to several percent \cite{STONE2}: $$ \frac{\tau (\Lambda _b)}{\tau (B_d)} \simeq 0.9 \; . \eqno(63b)$$ Similarly one estimates $\tau (\Xi _b)$ through a simple $1/m_Q^3$ scaling behaviour from the charm baryon lifetimes: $$\frac{\tau (\Xi _b^-)}{\tau (\Lambda _b)} \sim 1.1 \eqno(64a)$$ $$\frac{|\tau (\Xi _b^0)- \tau (\Lambda _b)|} {\tau (\Lambda _b)} < 0.1 \eqno(64b)$$ Obviously and not surprisingly there is some numerical fuzziness in these predictions; yet they seem to be unequivocal in stating that the $B_d$ lifetime exceeds the $\Lambda _b$ lifetime and the average beauty baryon lifetime by about 10 percent. However this prediction appears to be in serious (though not yet quite conclusive) disagreement with the data. If the predictions were based exclusively on adopting the quark model results for the baryonic expectation values, one could abandon them in a relatively light-hearted way: for it should not come as a shocking surprise that quark model results for baryonic matrix elements can be off-target by a substantial amount. Yet we have encountered a more serious problem here: data seem to contradict also the prediction based on an extrapolation from the observed lifetime pattern in the charm family; furthermore -- as discussed next -- the observed lifetime ratios of charm hadrons can be reproduced, within the expected uncertainties. This allows only one conclusion: if $\tau (B_d)$ indeed exceeds $\tau (\Lambda _b)$ by 25 - 30 \%, then a `theoretical price' has to be paid, namely that \noindent $\bullet$ the charm mass represents too low of a scale for allowing to go beyond merely qualitative predictions on charm baryon (or even meson) lifetimes, since it appears that corrections of order $1/m_c^4$ and higher are still important; \noindent $\bullet$ that the present agreement between theoretical expectations and data on charm baryon lifetimes is largely accidental and most likely would not survive in the face of more precise measurements! \noindent At the same time an intriguing puzzle arises: \noindent $\bullet$ Why are the quark model results for the relevant expectation values so much off the mark for beauty baryons? It is the deviation from unity in the lifetime ratios that is controlled by these matrix elements; finding a 30 \% difference rather than the expected 10 \% then represents a 300 \% error! Some new features emerge in $\Omega _b$ decays: since the $ss$-diquark system forms a spin-triplet, there are $1/m_b^2$ contributions to the semileptonic and nonleptonic $\Omega _b$ widths from the chromomagnetic operator. In order $1/m_b^3$ there arises a destructive interference in the $b\rightarrow c \bar cs$ channels. However a detailed discussion of $\Omega _b$ decays seems academic at the moment. \subsection{Semileptonic Branching Ratios of Beauty Hadrons} As briefly discussed before, one expects the semileptonic branching ratios for $B_d$ and $B_s$ mesons to practically coincide, in particular since semileptonic decays probe $\bar \tau (B_s)$, the average $B_s$ lifetime. It has already been pointed out that through order $1/m_b^2$ the expected value for $BR_{SL}(\Lambda _b)$ exceeds that for $BR_{SL}(B)$ by a few percent. In order $1/m_b^3$ the nonleptonic widths of both $B_d$ and $\Lambda _b$ states receive new contributions, with $\Gamma _{NL}(\Lambda _b)$ getting further enhanced relative to $\Gamma _{NL}(B)$ as expressed in eq.(63b). Thus one predicts $$BR_{SL}(\Lambda _b) < BR_{SL}(B_d) < BR_{SL}(B^-)$$ with the inequalities indicating differences of a few to several percent. If the present trend in the data persists, i.e. if the total $\Lambda _b$ width exceeds $\Gamma (B)$ by, say, 20 -25 \%, one would interprete this discrepancy as meaning that the nonleptonic -- but not the semileptonic -- width has received a unforeseen enhancement. In that case one expects $BR_{SL}(\Lambda _b)$ to fall below $BR_{SL}(B_d)$ by $\sim$ 20 \%. Putting all these observations together one concludes that the beauty lifetime averaged over $B_d$, $B^-$, $B_s$ and $\Lambda _b$ states should yield a value that is {\em smaller} -- by a few percent -- than $\frac{1}{2} \left[ BR_ {SL}(B_d) + BR_ {SL}(B^-)\right] \equiv \aver{BR_{SL}(B)}$: $$\aver{BR_{SL}(b)}< \aver{BR_{SL}(B)}$$ \section{Predictions on Charm Lifetimes} Considering the wealth of rather precise experimental information available on the lifetimes of charm hadrons one feels the urge to apply heavy quark expansions to charm decays as well. Yet in doing so one has to keep in mind that such a treatment might fail here for two basic reasons, one of which has just been stated: \noindent (i) The charm quark mass does not provide a sufficiently large scale to make the $1/m_c$ expansion converge quickly. To obtain an estimate of the size of the expansion parameter $\mu _{had}/m_c$ one can take the square root of the expression in eq.(35) representing the $1/m_c^2$ corrections: $$\frac{\mu _{had}}{m_c}\sim \sqrt{\frac{\aver {\mu _G^2}_D} {m_c^2}}\simeq 0.45 \, .\eqno(65)$$ This is not a small number although it is at least smaller than unity. \noindent (ii) As discussed in Sect.4 one has to go beyond {\em global} duality and invoke {\em local} duality to predict the decay widths of real hadrons from heavy quark expansions properly defined in Euclidean space. Yet a priori it is not clear at all whether contributions from `distant cuts' can be ignored since the charm quark mass does not exceed ordinary hadronic scales by a large amount. This concern is a posteriori strengthened by the following observation: Equating the observed semileptonic width of $D$ mesons with its theoretical expression through order $1/m_c^2$ (and assuming $|V(cs)| \simeq 1$) leads to the requirement $"m_c" \simeq 1.6$ GeV. This is however a high value relative to what is derived from charmonium spectroscopy, namely $m_c \leq 1.4$ GeV. A difference of $0.2$ GeV in $m_c$ might appear quite innocuous -- till one realizes that the corresponding semileptonic width depending on $m_c^5$ changes by a factor of two when $m_c$ is shifted by those $0.2$ GeV! At this point one might suspect that corrections of higher order in $1/m_c$ contribute {\em constructively} boosting the theoretical value. The analysis of ref.\cite{DIKEMAN} finds however that terms of order $1/m_c^3$ show a tendency to {\em decrease} $\Gamma _{SL}(D)$, and their inclusion does not help at all in reproducing the measured value of $\Gamma _{SL}(D)$ with $m_c\simeq 1.4$ GeV. This suggests that contributions from `distant cuts' which cannot be seen in any finite order of the $1/m_c$ expansion create this problem. For a different opinion on this point, see ref.\cite{CHERNYAK}. I draw the following somewhat tentative conclusions from these observations: \noindent $\bullet$ Predictions for the absolute size of charm hadron lifetimes cannot be trusted. \noindent $\bullet$ However it is quite conceivable that lifetime {\em ratios} do not suffer from such a fundamental uncertainty due to not-so-distant cuts. I will then explore the working hypothesis that the ratio of lifetimes can be trusted in principle -- though in practise only with a quite considerable grain of salt! The quark level Lagrangian is again well known. On the Cabibbo-allowed level there is now a single non-leptonic transition described by $${\cal L}_W^{\Delta C=1}(\mu = m_c)= \frac{4G_F}{\sqrt{2}} V_{cs}V_{ud}^* \{ c_1(\bar s_L\gamma _{\mu}c_L)(\bar u_L\gamma _{\mu}d_L) +c_2(\bar u_L\gamma _{\mu}c_L)(\bar s_L\gamma _{\mu}d_L) \} \eqno(66)$$ where the short-distance coefficients $c_{1,2}$ are now evaluated at a lowerJscale than in beauty decays yielding $$ c_1(LL+NLL)\simeq 1.32 \, , \; \; c_2(LL+NLL)\simeq - 0.58 \eqno(67)$$. The expressions for the semileptonic and non-leptonic decay widths of charm hadrons $H_c$ are quite analogous to the ones for beauty hadrons (though simpler since there is only one non-leptonic decay class rather than two); through order $1/m_c^2$ they are given by $$\frac{\Gamma _{SL,decay}(H_c)}{\Gamma _0}= \matel{H_c}{\bar cc}{H_c}_{norm}\cdot \left[ I_0(x,0,0) + \frac{\aver{\mu _G^2}_{H_c}}{m_c^2} (x\frac{d}{dx}-2)I_0(x,0,0)\right] \; , \eqno(68a)$$ $$\frac{\Gamma _{NL,decay}(H_c)}{\Gamma _0} = N_C\cdot \matel{H_c}{\bar cc}{H_c}_{norm}\cdot \left[A_0[I_0(x,0,0) + \frac{\aver{\mu _G^2}_{H_c}}{m_c^2} (x\frac{d}{dx}-2)I_0(x,0,0)] - \right.$$ $$\left. 8A_2 \frac{\aver{\mu _G^2}_{H_c}}{m_c^2} \cdot I_2(x,0,0,)\right]\; . \eqno(68b)$$ where now $$\Gamma _0 \equiv \frac{G_F^2m_c^5}{192 \pi ^3}|V(cs)|^2 \; , \; \; x=\frac{m_s^2}{m_c^2}\eqno(68c)$$ and the radiative corrections lumped into $A_0$ and $A_2$ are given by the appropriate values for $c_+$ and $c_-$, see eq. (67). With $x\sim 0.012$ one finds: $$I_0(x,0,0)|_{x=0.012}\simeq 0.91 \; \; , \; \; I_2(x,0,0)|_{x=0.012}\simeq 0.96 \; , $$ i.e. there is much less phase space suppression than for $b\rightarrow c$ transitions. \subsection{$D^+$ vs. $D^0$ Lifetimes} Analogously to the case of $\tau (B^-)$ vs. $\tau (B_d)$ the $D^+$ and $D^0$ lifetimes get first differentiated in order $1/m_c^3$ when PI and WA intervene. Both affect the lifetime ratio in the same direction, namely they enhance $\tau (D^+)$ over $\tau (D^0)$ with the {\em destructive interference} due to PI being the {\em dominant} effect. Quantitatively one finds through order $1/m_c^3$ by employing eqs.(48,49) with the appropriate substitutions and using $f_D \sim 200$ MeV: $$\frac{\tau (D^+)}{\tau (D^0)} \sim 2 \eqno(69)$$ For proper perspective one has to keep four observations in mind: \noindent (i) While the expectation expressed in eq.(69) does not coincide numerically with the measured value -- $\tau (D^+) /\tau (D^0) = 2.547 \pm 0.043$ -- it agrees with it to within $\sim 25$ \%. Such a deviation could be ascribed to $1/m_c^4$ contributions ignored in eq.(69). \noindent (ii) On the other hand, as discussed before, there is reason to doubt the reliability and thus validity of heavy quark expansions for treating nonperturbative corrections in {\em charm} decays. Yet I adopt, as already stated, the working hypothesis that a heavy quark expansion can be employed for treating the {\em ratios} of lifetimes, though not the lifetimes themselves. Such a conjecture is tested a posteriori by analysing the whole pattern of the expected and the observed lifetimes of the various charm hadrons. \noindent (iii) Even the perturbative corrections contain sizeable numerical uncertainties. To cite but one glaring example: the size and the nature of the socalled `hybrid' renormalization reflecting dynamics between the scales $m_c$ and $\mu _{had} \sim 0.5 - 1$ GeV is quite important quantitatively (they considerably enhance the destructive interference in $D^+$ decays.) Yet a leading-log treatment of those corrections seems woefully inadequate for dealing with such a small slice in momentum space. \noindent (iv) One should note also the following: while WA plays only a relatively minor role in inclusive rates (generating only a 10 \% or so difference in $D^+$ vs. $D^0$ lifetimes), it is likely to play a considerably more significant role in {\em exclusive} modes. \subsection{$D_s^+$ vs. $D^0$ Lifetimes} Since the impact of WA is reduced relative to that of PI in meson decays, it is natural to compare $\tau (D_s)$ to $\tau (D^0)$ rather than to $\tau (D^+)$. Such a comparison -- and of their semileptonic branching ratios -- touches upon several intriguing dynamical issues. A priori $\tau (D_s)$ and $\tau (D^0)$ could differ substantially from each other due to $SU(3)_{Fl}$ breaking and in particular due to a different weight of WA in the two cases. Yet the heavy quark expansion strongly suggests those two lifetimes to be close to each other. The analysis proceeds in several steps \cite{DS}. The two operators contributing in order $1/m_c^2$ are singlets under the light quark flavours; yet $SU(3)_{Fl}$ breaking could enter through their expectation values. For the chromomagnetic operator one has: $$\frac{1}{M_D}\matel{D^0}{\bar c i\sigma \cdot G c}{D^0} \simeq \frac{3}{2} (M^2_{D^{*0}}-M^2_{D^0})\eqno(70)$$ $$\frac{1}{M_{D_s}}\matel{D_s}{\bar c i\sigma \cdot G c}{D_s} \simeq \frac{3}{2} (M^2_{D_s^*}-M^2_{D_s})\eqno(71)$$ Since the measured values for the $D-D^*$ and the $D_s-D_s^*$ mass splittings are practically identical ( $M_{D^{*0}} - M_{D^0}=142.12 \pm 0.07$ MeV and $M_{D_s^*} - M_{D_s}=141.6 \pm 1.8$ MeV ), the chromomagnetic operator cannot induce an appreciable difference between $\tau (D^0)$ and $\tau (D_s)$. \footnote{The observation that the hyperfine splitting is largely independant of the flavour of the spectator (anti)quark can be understood intuitively in quark models.} Through order $1/m_c^3$ there are four distinct sources for a difference in $\Gamma (D_s)$ vs. $\Gamma (D^0)$ exceeding the 1\% level: (a) The decay $D_s\rightarrow \tau \nu$; (b) PI in those Cabibbo suppressed $D_s$ decays that are driven by the quark level transition $c\rightarrow s\bar su$; (c) the effects of $SU(3)_{Fl}$ breaking on the expectation values of the kinetic energy operator; (d) WA in nonleptonic $D^0$ and in nonleptonic as well as semileptonic $D_s$ decays. Corrections listed under (a), (b) and (d) are generated by dim-6 operators whereas the much less familiar correction (c) is derived from a dim-5 operator. \noindent (a) The width for the decay $D_s\rightarrow \tau \nu$ is completely determined in terms of the axial decay constant for the $D_s$ meson: $$\Gamma (D_s) = \frac{G_F^2m^2_{\tau}f^2_{D_s}M_{D_s}} {8\pi}|V(cs)|^2(1-m^2_{\tau}/M^2_{D_s})^2 . \eqno(72)$$ For $f_{D_s}\simeq 210$ MeV one gets $$\Gamma (D_s\rightarrow \tau \nu ) \simeq 0.03 \Gamma (D^0) . \eqno(73)$$ This effect necessarily {\em reduces} $\tau (D_s)$ relative to $\tau (D^0)$. \noindent (b) In $D_s$ decays PI appears in the Cabibbo-suppressed $c\rightarrow s \bar su$ channel. Its weight is expressed in terms of the matrix elements of the two four-fermion operators $$\matel{D_s}{(\bar c_L\gamma _{\mu}s_L) (\bar s_L\gamma _{\mu}c_L)}{D_s}, \; \; \matel{D_s}{(\bar c_L\gamma _{\mu}\lambda ^as_L) (\bar s_L\gamma _{\mu}\lambda ^ac_L)}{D_s} \eqno(74)$$ with known coefficients that are computed perturbatively (including the `hybrid' renormalization of these operators down from the scale $m_c$). The effect of PI is most reliably estimated from the observed difference in the $D^+$ and $D^0$ widths. It is easy to see that the structure of the operators responsible for PI in $D_s$ decays is exactly the same as in $D^+$ decays if one replaces the $d$ quark with the $s$ quark and adds the extra factor $\tan ^2\theta _C$; it is then {\em destructive} as well. Thus one arrives at: $$\delta \Gamma _{PI}(D_s) \simeq S\cdot \tan ^2\theta _C (\Gamma (D^+)-\Gamma (D^0)) \simeq -S\cdot 0.03 \Gamma (D^0)\eqno(75)$$ where the factor $S$ has been introduced to allow for $SU(3)_{Fl}$ breaking in the expectation values of the four-fermion operators. The factor $S$ is expected to exceed unity somewhat; in an factorization ansatz it is given by the ratio $f^2_{D_s}/f^2_D \simeq 1.4$, see eq.(39). Thus $$ \delta \Gamma _{PI}(D_s) \simeq - 0.04 \Gamma (D^0) \, . \eqno(76)$$ \noindent (c) The impact of the {\em chromomagnetic} operator on the $D_s-D^0$ lifetime ratio can be derived from the hyperfine splitting. Inserting the observed meson masses one obtains $$ \matel{D^0}{\bar c \frac{i}{2} \sigma \cdot G c}{D^0}_{norm} \simeq 0.413 \pm 0.002 \eqno(77a)$$ $$ \matel{D_s}{\bar c \frac{i}{2} \sigma \cdot G c}{D_s}_{norm} \simeq 0.433 \pm 0.006 \eqno(77b)$$ from which one can conclude: there is -- not surprisingly -- very little $SU(3)_{Fl}$ breaking in these matrix elements and a sizeable difference in $\tau (D_s)$ vs. $\tau (D^0)$ {\em cannot} arise from this source. \noindent The second dim-5 operator generating $1/m_c^2$ corrections is the kinetic energy operator $\bar c (i\vec D)^2c$ where $\vec D$ denotes the covariant derivative. Its expectation value describes the largely non-relativistic motion of the charm quark in the gluon background field inside the charm hadron. It reflects Lorentz time dilation and thus prolongs the {\em hadron} lifetime. On general grounds one expects it to extend $\tau (D_s)$ over $\tau (D^0)$, as seen as follows. The spatial wavefunction should be more concentrated around the origin for $D_s$ than for $D^0$ mesons. This implies, via the uncertainty principle, the mean value of $(\vec p_c)^2$ to be larger for $D_s$ than for $D^0$; in other words the charm quark undergoes more Fermi motion as constituent of $D_s$ than of $D$ mesons. The lifetime of the charm quark is then prolonged by time dilation to a higher degree inside $D_s$ than inside $D^0$ mesons. \noindent While the trend of this effect is quite transparent, its size is not yet clear. The relevant matrix elements $\matel{D_s}{\bar c (i\vec D)^2 c}{D_s}$ and $\matel{D}{\bar c (i\vec D)^2 c}{D}$ will be determined from QCD sum rules and lattice simulations in the future. Yet in the meantime one can estimate their size from the meson masses in the charm and beauty family according to the following prescription: $$ \matel{D_s}{\bar c (i\vec D)^2 c}{D_s} - \matel{D}{\bar c (i\vec D)^2 c}{D} \simeq \frac{2m_bm_c}{m_b-m_c} \{ [\aver{M_{D_s}}-\aver{M_D}]- [\aver{M_{B_s}}-\aver{M_B}]\} \eqno(78)$$ where as before $\aver{M_{D,D_s,B,B_s}}$ denote the spin averaged masses. Accordingly one finds $$\frac{\Delta \Gamma (D_s/D)}{\bar \Gamma}\equiv \frac{\Gamma _{Fermi}(D_s)- \Gamma _{Fermi}(D)} {\bar \Gamma } \simeq $$ $$ \simeq -\frac{m_b}{m_c(m_b-m_c)} \times \{ [\aver{M_{D_s}}-\aver{M_D}]- [\aver{M_{B_s}}-\aver{M_B}]\} \, .\eqno(79)$$ Since the $B_s^*$ mass has not been measured yet, one cannot give a specific numerical prediction and has to content oneself with semi-quantitative statements. A 10 MeV shift in any of the terms $\aver{M}$ corresponds numerically to the kinetic energy operator generating approximately a 1\% change in the ratio $\tau (D_s)/\tau (D^+)$. Invoking our present understanding of the heavy quark kinetic energy and its relationship to the hyperfine splitting one arrives at the following {\em conjecture}: $$ \frac{\Delta \Gamma (D_s/D)|_{(a)+(b)+(c)}}{\bar \Gamma} \sim {\cal O}(+\, few\, \% )\eqno(80)$$ \noindent (d) The mechanisms listed above in (a) - (c) taken together can be expected to extend the $D_s$ over the $D^0$ lifetime by at most a few percent. Comparing that estimate with the data allows us to conclude that WA cannot change $\tau (D_s)/\tau (D^0)$ by more than 10 \% . It is however quite unclear how to refine this estimate at present. For the quantitative impact of WA on charm meson lifetimes in general and on the ratio $\tau (D_s)/\tau (D^0)$ in particular is the most obscure theoretical element in the analysis. The uncertainty centers mainly on the question of how much the WA amplitude suffers from helicity suppression. \noindent In the valence quark approximation the answer is easily given to {\em lowest} order in the strong coupling: the WA amplitude is (helicity) suppressed by the ratio $m_q/m_c$, where $m_q$ denotes the largest quark mass in the final state. For a proper QCD treatment one has to use the {\em current} rather than the larger {\em constituent} mass, at least for the $1/m_c^3$ corrections. The WA amplitude is then small in $D^0$ decays where the helicity factor reads $m_s/m_c \sim 0.1$ and a fortiori in $D_s$ decays where only non-strange quarks are present in the final state. The emission of semi-hard gluons that is included by summing the leading log terms in the perturbative expansion cannot circumvent this suppression \cite{MIRAGE}. For such gluon corrections -- when properly accounted for -- drive the hybrid renormalization of the corresponding four-fermion operators; however they preserve the Lorentz structure of the lowest order term and therefore do not eliminate the helicity suppression. A helicity allowed amplitude arises in perturbation theory only at the subleading level of order $\alpha_s (m_c^2)/\pi$ and is thus expected to be numerically insignificant. \noindent On the other hand {\em nonperturbative} dynamics can quite naturally vitiate helicity suppression and thus provide the dominant source of WA. These nonperturbative effects enter through {\em non-factorizable} contributions to the hadronic matrix elements, as analyzed in considerable detail in refs.\cite{WA,DS}. As such we do not know (yet), how to predict their weight from first principles. However, as shown in ref.\cite{WA}, a detailed experimental study of the width of semileptonic decays and their lepton spectra -- in particular in the endpoint region -- in $D^0$ vs. $D_s$ and/or in $B^0$ vs. $B^+$ decays would allow us to extract the size of the matrix elements that control the weight of WA in all inclusive $B$ and $D$ decays. Before such data become available, we can draw only qualitative, or at best semi-quantitative conclusions: WA is not expected to affect the total lifetimes of $D_s$ and $D^0$ mesons by more than 10 - 20 \%, and their ratio by less. Furthermore WA does not necessarily decrease $\tau (D_s)/\tau (D^0)$; due to its reduced amplitude its leading impact on the lifetime might be due to its interference with the spectator amplitude and thus it might even {\em enhance} $\tau (D_s)/\tau (D^0)$! To summarize our findings on the $D_s-D^0$ lifetime ratio: (i) $SU(3)_{Fl}$ breaking in the expectation values of the dim-5 operators generating the leading nonperturbative corrections of order $1/m_c^2$ can -- due to `time dilatation' -- increase $\tau (D_s)$ by 3-5 \% over $\tau (D^0)$. (ii) On the $1/m_c^3$ level there arise three additional effects. Destructive interference in Cabibbo suppressed $D_s$ decays increases $\tau (D_s)$ again by 3-5 \%, whereas the single mode $D_s\rightarrow \tau \nu$ decreases it by 3 \%. The three phenomena listed so far combine to yield $\tau (D_s)/\tau (D^0) \simeq 1.0 - 1.07$. (iii) Any difference over and above that has to be attributed to WA. Therefore one has to interprete the measured $D_s-D^0$ lifetime ratio as more or less direct evidence for WA to contribute no more than 10-20 \% of the lifetime ratio between charm mesons. This is consistent with the indirect evidence discussed above that WA does not constitute the major effect there. (iv) These predictions can be refined in the future by two classes of more accurate data: analyzing the {\em difference} in the semileptonic spectra of charged and neutral mesons in the charm and beauty family allows to extract the size of the matrix elements that control the weight of WA; measuring the masses of $\Lambda _b$ and $B_s^*$ to better than 10 MeV allows to determine the expectation values of the kinetic energy operator. \subsection{Charm Baryon Lifetimes} The lifetimes of the weakly decaying charm baryons reflect a complex interplay of different dynamical mechanisms increasing or decreasing transition rates \cite{MARBELLA,BARYONS3}: $$\Gamma (\Lambda _c^+)= \Gamma _{decay}(\Lambda _c^+) + \Gamma _{WS}(\Lambda _c^+) - |\Delta \Gamma _{PI,-}(\Lambda _c)| \eqno(81a)$$ $$\Gamma (\Xi _c^0)= \Gamma _{decay}(\Xi _c^0) + \Gamma _{WS}(\Xi _c^0) + |\Delta \Gamma _{PI,+}(\Xi _c^0)| \eqno(81b)$$ $$\Gamma (\Xi _c^+)= \Gamma _{decay}(\Xi _c^+) + |\Delta \Gamma _{PI,+}(\Xi _c^+)| - |\Delta \Gamma _{PI,-}(\Xi _c^+)| \eqno(81c)$$ The explicit expressions for $\Gamma _{decay}$, $\Gamma _{WS}$ and $\Delta \Gamma _{PI,-}$ are obtained from eqs.(60) by the obvious substitutions; for $\Delta \Gamma _{PI,+}$ one finds $$\Delta \Gamma _{PI,+}(\Xi _c)\equiv \frac{G_F^2|V(cs)|^2m_c^2}{4\pi} c_+(2c_- + c_+) \matel{\Xi _c} {\bar c_L\gamma _{\mu}c\bar s_L\gamma _{\mu}s_L + \frac{2}{3}\bar c\gamma _{\mu}\gamma _5 c \bar s_L\gamma _{\mu}s_L} {\Xi _c}_{norm} \eqno(82)$$ On rather general grounds one then concludes: $$\tau (\Xi _c^0) < \tau (\Xi _c^+)\; , \; \; \; \tau (\Xi _c^0) < \tau (\Lambda _c^+) \eqno(83)$$ To go further requires calculating the relative weight of matrix elements of the four-quark operators. At present this can be done only by using quark wavefunctions as obtained from a potential ansatz (or from the quark model). In doing so one also has to pay proper attention to the normalization point appropriate for such an evaluation, i.e., one has to include ultraviolet as well as hybrid renormalization. One then arrives at the numbers quoted in Table 2. It has been argued \cite{MARBELLA} that in these expressions one should not use the `real' value for the decay constant, as expressed by $f_D$, and the observed mass splitting, but solely the leading contribution in a $1/m_Q^3$ expansion: $F_D \sim 400$ MeV and $M_{\Sigma ^+_c} - M_{\Lambda _c}\sim 400$ MeV. One reason for that is self-consistency since the widths have been calculated through order $1/m_Q^3$ only; the other -- and maybe the more telling one -- is based on the needs of phenomenology: for otherwise one cannot reproduce the observed {\em magnitude} of the lifetime differences. The predictions given in Table 2 agree remarkably well with the data within the experimental and theoretical uncertainties. \footnote{One should point out that those predictions were made before these data became available; data at that time suggested considerably larger lifetime ratios.} In the comparison of the $\Lambda _c$ and $D$ lifetimes one should note the following: WS and PI {\em counteract} each other in $\Gamma (\Lambda _c)$, see eq.(81a); nevertheless $\Gamma (\Lambda _c)\sim 2\cdot \Gamma (D)$ can be obtained. This results from three effects: (i) WA is very much reduced in $D$ decays. (ii) The {\em baryonic} expectation values of the four-quark operators are evaluated, as stated above, with the `static' decay constant, which is considerably larger than $f_D$ used for the {\em mesonic} expectation values. (iii) The $1/m_c^2$ contributions enhance $\Gamma (\Lambda _c)$ over $\Gamma (D)$, as briefly discussed below. Unfortunately there appears now a rather unpleasant `fly in the ointment': as discussed above, extrapolating from these prima facie successful predictions to the lifetimes of beauty baryons leads to a less-than-successful prediction on $\tau (\Lambda _b)/\tau (B_d)$. The decays of $\Omega _c$ baryons require -- and deserve -- a separate treatment: the $ss$ diquarks carry spin one and the resulting spin-spin interactions of $c$ with $ss$ lead to new effects. To be more specific: through order $1/m_c^3$ one has $$\Gamma (\Omega _c) = \Gamma _{decay}(\Omega _c) + |\Delta \Gamma _{PI,+}(\Omega _c)| \eqno(84)$$ with both quantities on the right-hand-side of eq.(84) differing from the corresponding ones for $\Lambda _c$ or $\Xi _c$ decays. (i) Firstly, $\matel{\Omega _c}{\bar c i\sigma \cdot G c}{\Omega _c}\neq 0 = \matel{\Lambda _c}{\bar c i\sigma \cdot G c}{\Lambda _c}$, see eqs.(17b,17c,35). Secondly, it is quite conceivable that these spin-spin interactions can create an appreciable difference in the kinetic energy of the $c$ quark inside $\Omega _c$ and $\Lambda _c$ baryons. Thus there arise differences of order $1/m_c^2$ in the total as well as in the semileptonic widths of $\Omega _c$ and $\Lambda _c$ baryons: $$\Gamma _{decay}(\Omega _c) \neq \Gamma _{decay}(\Lambda _c) \; \; , \; \; \Gamma _{SL,decay}(\Omega _c) \neq \Gamma _{SL,decay}(\Lambda _c) \eqno(85)$$ (ii) These spin-spin interactions also affect the expectation values of the dimension-six four-fermion operators that control the strength of WS and the interference effects, see eqs.(24,25), in order $1/m_c^3$. \noindent Taking everything together one estimates the $\Omega _c$ to be the shortest lived charm baryon, as shown in Table 2. This is quite remarkable since it means that the constructive interference in $\Omega _c$ decays outweighs the {\em combined} effect of constructive interference and WS in $\Xi _c^0$ decays. It is actually not unexpected on intuitive grounds: in $\Omega _c$ transitions interference can happen with {\em both} constituent $s$ quarks whereas in $\Xi _c$ baryons there is only one such $s$ quark. This shows that inclusive weak decay rates can be harnessed to probe the internal structure of charm baryons in a novel way. \subsection{Semileptonic Branching Ratios of Charm Hadrons} As expected, the ratio of the semileptonic $D^+$ and $D^0$ branching ratios agrees, within the errors, with the ratio of their lifetimes: $$2.23 \pm 0.42= \frac{BR_{SL}(D^+)}{BR_{SL}(D^0)}\simeq \frac{\tau (D^+)}{\tau (D^0)}=\, 2.547\, \pm \, 0.043 \eqno(86)$$ Also the {\em absolute} size of, say, $BR_{SL}(D^0)$ is of interest. If the $D^+ - D^0$ lifetime difference is generated mainly by a destructive interference affecting $D^+$ transitions, then $D^0$ decays should proceed in a largely normal way and the semileptonic branching ratio of $D^0$ mesons should be at its `normal' value. In a parton model description one finds $BR(D\rightarrow l\nu X_s) \equiv BR(c\rightarrow l\nu s) \sim 16\%$. If that number represented the proper yardstick for normal decays, one would actually conclude that $D^+$ decays more or less normally, whereas the nonleptonic decays of $D^0$ are strongly enhanced; i.e., it would imply that it is actually WA that provides the dominant mechanism for the lifetime difference. This long-standing `fly in the ointment' can be removed now. For at order $1/m_Q^2$ the chromomagnetic operator $\bar Qi\sigma \cdot G Q$ generates an isoscalar enhancement in the nonleptonic widths of the charged and neutral mesons \cite{BUV}. It turns out that the corresponding reduction in the semileptonic branching ratios of $B$ mesons is rather insignificant, namely $\sim$ 2 \% . Yet for charm mesons this reduction is quite large, namely around 40 \%, since it is scaled up by $(m_b/m_c)^2$ and less reduced by colour factors. The semileptonic branching ratio for $D$ mesons is thus around 9 \% through order $1/m_c^2$ -- before the dimension-six four-fermion operators reduce $\Gamma _{NL}(D^+)$ by a factor of almost two and enhance $\Gamma _{NL}(D^0)$ by a moderate amount. As discussed above, we have to take these numbers for charm transitions with a grain of salt; on the other hand the emerging pattern in the ratio of semileptonic branching ratios as well as their absolute size is self-consistent! It should be apparent from the discussion on $\tau (D_s)$ vs. $\tau (D^0)$ that the semileptonic widths and branching ratios of $D_s$ and $D^0$ mesons practically coincide through order $1/m_c^2$. In order $1/m_c^3$ WA can contribute to $\Gamma _{SL}(D_s)$ (as well as to $\Gamma _{NL}(D_s)$ and $\Gamma _{NL}(D^0)$). Yet from $\tau (D_s)\simeq \tau (D^0)$ one infers that the relative weight of WA is quite reduced; therefore one expects $BR_{SL}(D_s)$ to differ from $BR_{SL}(D^0)$ by not more than 10\% or so in {\em either} direction: $$1\, \pm \sim {\rm few}\% = \frac{\tau (D_s)}{\tau (D^0)} \simeq \frac{BR_{SL}(D_s)}{BR_{SL}(D^0)}=1\, \pm \sim 10\% \eqno(87)$$ The semileptonic width is not universal for all charm hadrons due to ${\cal O}(1/m_c^2)$ contributions even when one ignores Cabibbo suppressed channels. One actually estimates $$ \Gamma _{SL}(D) / \Gamma _{SL}(\Lambda _c) / \Gamma _{SL}(\Omega _c) \, \sim \, 1/1.5/1.2 \, .\eqno(88)$$ Therefore the ratio of semileptonic branching ratios of mesons and baryons does not faithfully reproduce the ratio of their lifetimes. Instead one expects $$BR_{SL}(\Lambda _c) > BR_{SL}(D^0)\times \tau (\Lambda _c)/\tau (D^0) \simeq 0.5\, BR_{SL}(D^0) \, . \eqno(89)$$ \section{What Have We Learned and What Will We Learn?} Second generation theoretical technologies have been developed for treating heavy-flavour decays that are directly related to QCD without the need for invoking a `deus ex machina'. As far as {\em inclusive} heavy-flavour decays are concerned, the relevant technology is based on an operator expansion in inverse powers of the heavy quark mass \footnote{To repeat it one last time: This is {\em not} HQET!}. It allows to express the leading nonperturbative corrections through the expectation values of a small number of dimension-five and -six operators. Basically all such matrix elements relevant for {\em meson} decays can reliably be related to other observables; this allows to extract their size in a model-independant way. Baryons, however, possess a more complex internal structure, which becomes relevant for their decays in order $1/m_Q^3$. At present one has to rely on quark model calculations to determine the expectation values of the dimension-six operators relevant for lifetime differences. The numerical results of such computations are of dubious reliability; accordingly predictions for lifetime ratios involving heavy-flavour {\em baryons} suffer from larger uncertainties than those involving only mesons. There is also a positive side to this, though: measuring the lifetimes (and if at all possible also the semileptonic branching ratios) of the various heavy-flavour baryons sheds new light onto their internal strong dynamics. Comprehensive studies of inclusive weak decays in the charm baryon family are particularly promising from a practical point for such studies on hadronic structure; for the pre-asymptotic corrections are much larger here than for beauty baryons. At present there exists one rather glaring phenomenological problem for the heavy quark expansion and -- not surprisingly, as just indicated -- it concerns baryon decays: the observed $\Lambda _b$ lifetime is clearly shorter than predicted relative to the $B_d$ lifetime. Unless future measurements move it up significantly, one has to pay a theoretical price for that failure. As stated above, for the time being one has to employ a quark model to determine the size of the baryonic expectation values of the relevelant dimenion-six operators. It is not surprising per se that quark model computations can yield numerically incorrect results, in particular when sizeable cancellations occur between different contributions. The more serious problem is provided by the following aspects: the analoguous treatment of charm baryons, using the same quark models, had yielded predictions that give not only the correct pattern, but also numerical values for the lifetime ratios which are surprisingly close to the data considering the uncertainties due to higher order terms in the $1/m_c$ expansion. Furthermore scaling these charm lifetime ratios up to the beauty scale, using a $1/m_Q^2$ and $1/m_Q^3$ behaviour for the pre-asymptotic corrections from the dimension-five and -six operators, respectively, also tells us that the $\Lambda _b$ lifetime should be shorter than the $B_d$ lifetime by no more than 10 \%. To the degree that the observed value for $\tau (\Lambda _b)/\tau (B_d)$ falls below 0.9 one has to draw the following conclusion: one cannot trust the numerical results of quark model calculations for {\em baryonic} matrix elements -- not very surprising by itself; yet furthermore and more seriously it would mean that $1/m_Q^4$ or even higher order contributions are still relevant in charm baryon decays before fading away for beauty decays. Then one had to view the apparently successful predictions on the lifetimes of charm baryons as largely coincidental -- a quite sobering result! A {\em future} discrepancy between the predictions on $\tau (B^+)/\tau (B_d)$ or $\tau (B_d)/\bar \tau (B_s)$ and the data \footnote{It has to be stressed again that $\bar \tau (B_s)$ refers to the algebraic average of the two $B_s$ lifetimes.} -- in particular an observation that the lifetime for $B^+$ mesons is definitely {\em shorter} than for $B_d$ mesons -- would have quite fundamental consequences. For the leading deviation of these ratios from unity arises at order $1/m_b^3$ and should provide a good approximation since the expansion parameter is small: $\mu _{had}/m_b \sim 0.1$. The size of this term is given by the expectation value of a four-fermion operator expressed in terms of $f_B$. A failure in this simple situation would raise very serious doubts about the validity or at least the practical relevance of the $1/m_Q$ expansion for treating even fully inclusive nonleptonic transitions; this would leave only semileptonic transitions in the domain of their applicability. Such a breakdown of quark-hadron duality would {\em a priori} appear as a quite conceivable and merely disappointing scenario. However such an outcome would have to be seen as quite puzzling {\em a posteriori}; for in our analysis we have not discerned any sign indicating the existence of such a fundamental problem or a qualitative distinction between nonleptonic and semileptonic decays \cite{BLOKMANNEL,OPTICAL}. Thus even a failure would teach us a valuable, albeit sad lesson about the intricacies of the strong interactions; for the heavy quark expansion is directly and unequivocally based on QCD with the only additional assumption concerning the workings of quark-hadron duality! The theoretical analysis of the lifetimes of heavy-flavour hadrons can be improved, refined and extended: \noindent $\bullet$ {\em improved} by a better understanding of quark-hadron duality, preferably by deriving it from QCD or at least by analysing how it operates in the lepton and photon spectra of semileptonic and radiative $B$ decays, respectively; \noindent $\bullet$ {\em refined} by a reliable determination of in particular, but not only, the baryonic expectation values of the relevant dimension-six operators; \noindent $\bullet$ {\em extended} by treating $\Xi _b$ (and $\Omega _b$) decays. The extension will certainly have been made very soon; the refinement should be obtained in the foreseeable future, probably from a lattice simulation of QCD; when finally quark-hadron duality will be derived from QCD cannot be predicted, of course. There is a corresponding list of future measurements that are most likely to probe and advance our understanding: \noindent $\bullet$ measure $\tau (\Lambda _b)$, $\tau (\Xi ^- _b)$ and $\tau (\Xi ^0 _b)$ {\em separately} with good accuracy; \noindent $\bullet$ confirm $\tau (B_d)\simeq \bar \tau (B_s)$ within an accuracy of very few percent; \noindent $\bullet$ verify that $\tau (B^+)$ {\em exceeds} $\tau (B_d)$ by a few to several percent; \noindent $\bullet$ in charm decays determine the lifetimes of $\Xi _c^{0,+}$ and $\Omega _c$ baryons with a few percent accuracy and measure $\tau (D_s)/\tau (D^0)$ with 1\% precision. \noindent {\bf Acknowledgements:} I am deeply indebted to my collaborators B. Blok, M. Shifman, N. Uraltsev and A. Vainshtein for sharing their insights with me. Prodding from many experimental colleagues, in particular S. Paul and V. Sharma, has been instrumental in making me face reality (I think). I have been fortified by many excellent chinese dinners prepared late at night by Bai-Ju's in South Bend.
\section{Introduction} After the celebrated discovery of the top quark \cite{abe}, the Higgs boson is the last missing link in the Standard Model (SM). The detection of this particle and the study of its characteristics are among the prime objectives of present and future high-energy colliding-beam experiments. Following Bjorken's proposal \cite{bjo}, the Higgs boson is currently being searched for with the CERN Large Electron-Positron Collider (LEP1) and the SLAC Linear Collider (SLC) via $e^+e^-\to Z\to f\bar fH$. At the present time, the failure of this search allows one to rule out the mass range $M_H\le64.3$~GeV at the 95\% confidence level \cite{jan}. The quest for the Higgs boson will be continued with LEP2 by exploiting the Higgs-strahlung mechanism \cite{ell,iof}, $e^+e^-\to ZH\to f\bar fH$. In next-generation $e^+e^-$ linear supercolliders (NLC), also $e^+e^-\to\bar\nu_e\nu_eH$ via $W^+W^-$ fusion and, to a lesser extent, $e^+e^-\to e^+e^-H$ via $ZZ$ fusion will provide copious sources of Higgs bosons. Once a novel scalar particle is discovered, it will be crucial to decide if it is the very Higgs boson of the SM or if it lives in some more extended Higgs sector. To that end, precise knowledge of the SM predictions will be mandatory, {\it i.e.}, quantum corrections must be taken into account. The status of the radiative corrections to the production and decay processes of the SM Higgs boson has recently been summarized \cite{pr}. Since the top quark is by far the heaviest established elementary particle, with a pole mass of $M_t=(180\pm12)$~GeV \cite{abe}, the leading high-$M_t$ terms, of ${\cal O}(G_FM_t^2)$, are particularly important, and it is desirable to acquire information on their quantumchromodynamical (QCD) corrections. During the last year, a number of papers have appeared in which the two-loop ${\cal O}(\alpha_sG_FM_t^2)$ corrections to various Higgs-boson production and decay processes are presented. These processes include $H\to f\bar f$, with $f\ne b$ \cite{hll} and $f=b$ \cite{ks1,kwi}, $Z\to f\bar fH$ and $e^+e^-\to ZH$ \cite{ks2}, $e^+e^-\to\bar\nu_e\nu_eH$ via $W^+W^-$ fusion \cite{ks3}, $gg\to H$ \cite{ks3,gam}, and more \cite{ks3}. In this paper, we shall take the next step and tackle with three-loop ${\cal O}(\alpha_s^2G_FM_t^2)$ corrections. To keep matters as simple as possible, we shall restrict our considerations to light Higgs bosons, with $M_H\ll M_t$, and to reactions with colourless particles in the initial and final states. Such reactions typically involve the $\ell^+\ell^-H$, $W^+W^-H$, and $ZZH$ couplings together with gauge couplings of the $W$ and $Z$ bosons to leptons. We are thus led to incorporate the next-to-leading QCD corrections in the low-$M_H$ effective $\ell^+\ell^-H$, $W^+W^-H$, and $ZZH$ interaction Lagrangians. This will be achieved in Section~2. Recently, the ${\cal O}(\alpha_s^2G_FM_t^2)$ correction to $\Delta\rho$ has been calculated and found to be sizeable \cite{avd}. This is relevant for present and future precision tests of the standard electroweak theory. It is of great theoretical interest to find out whether the occurrence of significant ${\cal O}(\alpha_s^2G_FM_t^2)$ corrections is specific to $\Delta\rho$ or whether this is a common feature in the class of electroweak observables with a quadratic $M_t$ dependence at one loop. In the latter case, there must be some underlying principle which is able to explain this phenomenon. Our analysis will put us into a position where we can investigate this matter for four independent quantities. We shall return to this issue in Section~5. The complete evaluation of the one-loop electroweak correction to a process which involves more than four external particles is enormously intricate. To our knowledge, the literature does not contain a single example of such a calculation. However, the so-called improved Born approximation (IBA) \cite{iba} allows us to conveniently extract at least the dominant fermionic loop corrections. As a by-product of our analysis, we shall illustrate the usefulness of the IBA for Higgs-boson production and decay in high-energy $e^+e^-$ collisions. The appropriate formalism will be developed in Section~3. This paper is organized as follows. In Section~2, we shall extend the low-$M_H$ effective $\ell^+\ell^-H$, $W^+W^-H$, and $ZZH$ interaction Lagrangians to ${\cal O}(\alpha_s^2G_FM_t^2)$. In the $G_F$ formulation of the electroweak on-shell scheme, knowledge of the QCD-corrected $W^+W^-H$ coupling is sufficient to control the related four- and five-point Higgs-boson production and decay processes which emerge by connecting one or both of the $W$ bosons with lepton lines, respectively. Contrariwise, the corresponding processes involving a $ZZH$ coupling receive additional QCD corrections from the gauge sector, which we shall evaluate by invoking the IBA in Section~3. In Section~4, we shall quantitatively analyze the phenomenological consequences of our results. Section~5 contains our conclusions. \section{Effective Lagrangians} Throughout this paper, we shall work in the electroweak on-shell renormalization scheme \cite{aok}, with $G_F$ as a basic parameter, and define $c_w^2=1-s_w^2=M_W^2/M_Z^2$ \cite{sir}. In particular, this implies that the lowest-order formulae are expressed in terms of $G_F$, $c_w$, $s_w$, and the physical particle masses. The self-energies of the $W$, $Z$, and Higgs bosons to ${\cal O}(\alpha_s^2G_FM_t^2)$ for zero external four-momentum squared will be the basic ingredients of our analysis. While the results for the $W$ and $Z$ bosons are now well established \cite{avd}, the Higgs-boson self-energy requires a separate analysis, which will be performed here. Our calculation will proceed along the lines of Ref.~\cite{avd}. We shall employ dimensional regularization in $n=4-2\epsilon$ space-time dimensions and introduce a 't~Hooft mass, $\mu$, to keep the coupling constants dimensionless. We shall suppress terms containing $\gamma_E-\ln(4\pi)$, where $\gamma_E$ is Euler's constant. These terms may be retrieved by substituting $\mu^2\to4\pi e^{-\gamma_E}\mu^2$. In the modified minimal-subtraction ($\overline{\mbox{MS}}$) scheme \cite{msb}, these terms are subtracted along with the poles in $\epsilon$. This is also true for the relation between the $\overline{\mbox{MS}}$ and pole masses of the quarks, so that these terms are also absent when the quark masses are renormalized according to the on-shell scheme. Since we wish to extract the leading high-$M_t$ terms, we may neglect the masses of all virtual particles, except for the top quark. As usual, we shall take $\gamma_5$ to be anticommuting for $n$ arbitrary. We shall choose a covariant gauge with an arbitrary gauge parameter for the gluon propagator. This will allow us to explicitly check that our final results are gauge independent. The requirement that the expressions for physical observables be renormalization-group (RG) invariant will serve as a further check for our calculation. Large intermediate expressions will be treated with the help of FORM 2.0 \cite{ver}. The tadpole integrals which enter the one- and two-loop calculations may be solved straightforwardly, even for arbitrary powers of propagators. The three-loop case is more involved. After evaluating the traces, the scalar integrals may be reduced by decomposing the scalar products in the numerator into appropriate combinations of the factors in the denominator. Subsequently, recurrence relations derived using the integration-by-parts method \cite{che} may be applied to reduce any scalar Feynman integral to a small number of so-called master diagrams, which remain to be calculated by hand. More technical details may be found in Ref.~\cite{avd}. Prior to listing our results, we shall introduce our notation. We take the colour gauge group to be SU($N_c$); $C_F=(N_c^2-1)/(2N_c)$ and $C_A=N_c$ are the Casimir operators of its fundamental and adjoint representations, respectively. As is usually done for SU($N_c$), we fix the trace normalization of the fundamental representation to be $T_F=1/2$. In our numerical analysis, we set $N_c=3$. We explicitly include five massless quark flavours plus the massive top quark in our calculation, so that we have $n_f=6$ active quark flavours altogether, {\it i.e.}, we must not consider $n_f$ as a free parameter. We denote the QCD renormalization scale by $\mu$. We evaluate the strong coupling constant, $\alpha_s(\mu)$, at next-to-leading order (two loops) in the $\overline{\mbox{MS}}$ scheme, from \begin{equation} \label{as} {\alpha_s(\mu)\over\pi}={1\over\beta_0\ln(\mu^2/\Lambda_{\overline{MS}}^2)} \left[1-{\beta_1\over\beta_0^2}\,{\ln\ln(\mu^2/\Lambda_{\overline{MS}}^2) \over\ln(\mu^2/\Lambda_{\overline{MS}}^2)}\right], \end{equation} where $\Lambda_{\overline{MS}}$ is the asymptotic scale parameter appropriate for $n_f=6$ and \cite{wil} \begin{eqnarray} \label{beta} \beta_0&\hspace*{-2.5mm}=\hspace*{-2.5mm}&{1\over4}\left({11\over3}C_A-{2\over3}n_f\right)={7\over4}, \nonumber\\ \beta_1&\hspace*{-2.5mm}=\hspace*{-2.5mm}&{1\over16}\left({34\over3}C_A^2-2C_Fn_f-{10\over3}C_An_f\right) ={13\over8} \end{eqnarray} are the first two coefficients of the Callan-Symanzik beta function of QCD. We define $a=4h=\alpha_s(\mu)/\pi$, $x_t=[G_Fm_t^2(\mu)/8\pi^2\sqrt2]$, $X_t=(G_FM_t^2/8\pi^2\sqrt2)$, $l=\ln[\mu^2/m_t^2(\mu)]$, and $L=\ln(\mu^2/M_t^2)$, where $m_t(\mu)$ and $M_t$ are the $\overline{\mbox{MS}}$ and pole masses of the top quark, respectively, and $G_F$ is Fermi's constant. Using the two-loop relation between $m_t(M_t)$ and $M_t$ \cite{gra} along with the RG equation for $m_t(\mu)$, we find \begin{eqnarray} \label{mass} {m_t(\mu)\over M_t}&\hspace*{-2.5mm}=\hspace*{-2.5mm}&1+hC_F(-3L-4) +h^2C_F\left\{L^2\left({9\over2}C_F-{11\over2}C_A+n_f\right) +L\left({21\over2}C_F-{185\over6}C_A \right.\right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left. {13\over3}n_f\right)-12\zeta(2)+6 +C_F\left[-12\zeta(3)+6\zeta(2)(8\ln2-5)+{7\over8}\right] \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left. C_A\left[6\zeta(3)+8\zeta(2)(-3\ln2+1)-{1111\over24}\right] +n_f\left[4\zeta(2)+{71\over12}\right]\right\} \nonumber\\ &\hspace*{-2.5mm}\approx\hspace*{-2.5mm}&1-a\left(L+{4\over3}\right) -a^2\left({3\over8}L^2+{35\over8}L+9.125\,451\right). \end{eqnarray} Riemann's zeta function takes on the values $\zeta(2)=\pi^2/6$, $\zeta(3)\approx1.202\,057$, and $\zeta(4)=\pi^4/90$. The numerical constants \cite{avd} \begin{eqnarray} S_2&\hspace*{-2.5mm}=\hspace*{-2.5mm}&{4\over9\sqrt3}\mathop{{\mbox{Cl}}_2}\nolimits\left({\pi\over3}\right) \approx0.260\,434,\nonumber\\ D_3&\hspace*{-2.5mm}\approx\hspace*{-2.5mm}&-3.027\,009,\nonumber\\ B_4&\hspace*{-2.5mm}=\hspace*{-2.5mm}&16\mathop{{\mbox{Li}}_4}\nolimits\left({1\over2}\right)-{13\over2}\zeta(4)-4\zeta(2)\ln^22 +{2\over3}\ln^42 \approx-1.762\,800, \end{eqnarray} where $\mathop{{\mbox{Cl}}_2}\nolimits{}$ is Clausen's function and $\mathop{{\mbox{Li}}_4}\nolimits{}$ is the quadrilogarithm, occur in the evaluation of the three-loop master diagrams. In the following , we shall frequently make use of the QCD expansion of $\Delta\rho$ through ${\cal O}(\alpha_s^2G_FM_t^2)$. {}For the reader's convenience, we shall list it here for $N_c=3$ and $n_f=6$. The $\overline{\mbox{MS}}$ and on-shell results read \cite{avd} \begin{eqnarray} \Delta\bar\rho&\hspace*{-2.5mm}\approx\hspace*{-2.5mm}& N_cx_t\left[1+a(2\,l-0.193\,245) +a^2\left({15\over4}l^2+2.025\,330\,l-3.969\,560\right)\right], \\ \label{drhoos} \Delta\rho&\hspace*{-2.5mm}\approx\hspace*{-2.5mm}& N_cX_t[1-2.859\,912\,a-a^2(5.004\,846\,L+14.594\,028)], \end{eqnarray} respectively. To start with, we shall construct the low-$M_H$ effective $\ell^+\ell^-H$ interaction Lagrangian through ${\cal O}(\alpha_s^2G_FM_t^2)$. In the following, bare quantities will be labelled with the superscript 0. The bare $\ell^+\ell^-H$ Lagrangian reads \begin{equation} \label{llh} {\cal L}_{\ell\ell H}=-m_\ell^0\bar\ell^0\ell^0{H^0\over v^0}, \end{equation} where $v$ denotes the Higgs vacuum expectation value. The renormalizations of the lepton mass and wave function do not receive corrections in ${\cal O}(\alpha_s^nG_FM_t^2)$, where $n=0,1,2$, so that we may replace $m_\ell^0$ and $\ell^0$ with their renormalized counterparts. In the $G_F$ formulation of the on-shell scheme, we have \cite{hff} \begin{equation} \label{du} {H^0\over v^0}=2^{1/4}G_F^{1/2}H(1+\delta_u), \end{equation} with \begin{equation} \delta_u=-{1\over2}\left[{\Pi_{WW}(0)\over M_W^2}+\Pi_{HH}^\prime(0)\right]. \end{equation} Here, $\Pi_{WW}(q^2)$ and $\Pi_{HH}(q^2)$ are the $W$- and Higgs-boson self-energies for external momentum $q$, respectively, and the subscript $u$ is to remind us that this term appears as a universal building block in the radiative corrections to all production and decay processes of the Higgs boson. Consequently, the renormalized version of Eq.~(\ref{llh}) reads \begin{equation} \label{lllh} {\cal L}_{\ell\ell H}=-2^{1/4}G_F^{1/2}m_\ell\bar\ell\ell H(1+\delta_u). \end{equation} The one-loop expressions for $\Pi_{WW}(q^2)$ and $\Pi_{HH}(q^2)$ have been presented in Ref.~\cite{hzz}. The leading-order QCD corrections to $\Pi_{WW}(q^2)$ and $\Pi_{HH}(q^2)$ for arbitrary quark masses have been found in Refs.~\cite{djo,hll}, respectively. The ${\cal O}(\alpha_sG_FM_t^2)$ term of $\delta_u$ has independently been obtained in Ref.~\cite{kwi} by using the computational technique outlined above at the two-loop level. Here, we shall extend this analysis to ${\cal O}(\alpha_s^2G_FM_t^2)$. The ${\cal O}(\alpha_s^2G_FM_t^2)$ term of $\Pi_{WW}(0)$ may be found in Ref.~\cite{avd}. The Feynman diagrams pertinent to $\Pi_{HH}(q^2)$ in ${\cal O}(\alpha_s^2G_FM_t^2)$ come in twenty different topologies. Typical examples are depicted in Fig.~\ref{one}. We shall renormalize the strong coupling constant and the top-quark mass according to the $\overline{\mbox{MS}}$ scheme. The appropriate counterterms are listed in Ref.~\cite{gra}. In this way, we obtain \begin{eqnarray} \label{pihh} \Pi_{HH}^\prime(0)&\hspace*{-2.5mm}=\hspace*{-2.5mm}&N_cx_t\left\{{2\over\epsilon}+2l-{4\over3} +hC_F\left(-{6\over\epsilon^2}+{5\over\epsilon}+6l^2-10l-{37\over6}\right) +h^2C_F\left[27\zeta(3)+6 \right.\right. \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_F\left({12\over\epsilon^3}-{12\over\epsilon^2} +{1\over\epsilon}\left(24\zeta(3)-{119\over6}\right) +l\left(72\zeta(3)-{93\over2}\right)+24B_4-108\zeta(4) \right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left. 106\zeta(3)+{331\over12}\right) +C_A\left({22\over3\epsilon^3}-{83\over3\epsilon^2} +{1\over\epsilon}\left(-12\zeta(3)+{77\over3}\right) +{22\over3}l^3 \right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left. 14l^2+l\left(-36\zeta(3)-{961\over18}\right)-12B_4+54\zeta(4) -{55\over3}\zeta(3)-7\right) \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left.\left. n_f\left(-{4\over3\epsilon^3}+{10\over3\epsilon^2}-{8\over3\epsilon} -{4\over3}l^3+{65\over9}l-{32\over3}\zeta(3)-3\right)\right]\right\}. \end{eqnarray} When we combine Eq.~(\ref{pihh}) with the corresponding expression for $\Pi_{WW}(0)$ \cite{avd}, the ultraviolet divergences cancel, and we obtain \begin{eqnarray} \bar\delta_u&\hspace*{-2.5mm}=\hspace*{-2.5mm}&N_cx_t\left\{{7\over6} +hC_F\left(7l-2\zeta(2)+{19\over3}\right) +h^2C_F\left[243S_2-{449\over6}\zeta(3)-{14\over3}\zeta(2)+{79\over3} \right.\right. \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_F\left(21l^2+l\left(-12\zeta(2)-{1\over2}\right)+4B_4+2D_3 -{1053\over2}S_2+2\zeta(4)+{599\over3}\zeta(3)-{259\over9}\zeta(2) \right.\nonumber\\ &\hspace*{-2.5mm}-\hspace*{-2.5mm}&\left. {3043\over72}\right) +C_A\left({77\over6}l^2+l\left(-{22\over3}\zeta(2)+{1097\over18}\right) -2B_4-D_3+{1053\over4}S_2+15\zeta(4) \right.\nonumber\\ &\hspace*{-2.5mm}-\hspace*{-2.5mm}&\left.\left.\left.\!\! {509\over6}\zeta(3) -{73\over3}\zeta(2) +{953\over24}\right) +n_f\left(\!-{7\over3}l^2+l\left({4\over3}\zeta(2)-{73\over9}\right) \!-{8\over3}\zeta(3)+{14\over3}\zeta(2)-{55\over12}\right)\right]\right\} \nonumber\\ &\hspace*{-2.5mm}\approx\hspace*{-2.5mm}& {7\over6}N_cx_t\left[1+a(2\,l+0.869\,561) +a^2\left({15\over4}l^2+6.010\,856\,l-2.742\,226\right)\right]. \end{eqnarray} With the help of Eq.~(\ref{mass}), we may eliminate $m_t(\mu)$ in favour of $M_t$, which leads to \begin{eqnarray} \label{duos} \delta_u&\hspace*{-2.5mm}=\hspace*{-2.5mm}&N_cX_t\left\{{7\over6} +hC_F\left(-2\zeta(2)-3\right) +h^2C_F\left[243S_2-{449\over6}\zeta(3)-{98\over3}\zeta(2)+{121\over3} \right.\right. \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_F\left(4B_4+2D_3-{1053\over2}S_2+2\zeta(4)+{515\over3}\zeta(3) +\zeta(2)\left(112\ln2-{745\over9}\right)-{146\over9}\right) \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_A\left(L\left(-{22\over3}\zeta(2)-11\right) -2B_4-D_3+{1053\over4}S_2+15\zeta(4)-{425\over6}\zeta(3) \right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left.\left.\left. \zeta(2)\left(-56\ln2-{17\over3}\right)-{2459\over36}\right) +n_f\left(L\left({4\over3}\zeta(2)+2\right) -{8\over3}\zeta(3)+14\zeta(2)+{83\over9}\right)\right]\right\} \nonumber\\ &\hspace*{-2.5mm}\approx\hspace*{-2.5mm}&{7\over6}N_cX_t[1-1.797\,105\,a-a^2(3.144\,934\,L+16.200\,847)]. \end{eqnarray} Equation~(\ref{duos}) reproduces the ${\cal O}(G_FM_t^2)$ and ${\cal O}(\alpha_sG_FM_t^2)$ terms found in Refs.~\cite{hff,hll}, respectively. We observe that the new ${\cal O}(\alpha_s^2G_FM_t^2)$ term in Eq.~(\ref{duos}) enhances the QCD correction and thus supports the screening of the leading-order $M_t$ dependence. The choice $\mu=M_t$ is singled out, since it eliminates the terms containing $L$ in Eq.~(\ref{duos}). The nonlogarithmic coefficient of $(\alpha_s/\pi)^2$ in Eq.~(\ref{duos}) is relatively large; it exceeds the corresponding coefficient of $\Delta\rho$ in Eq.~(\ref{drhoos}) by approximately 11\%. If we consider the ratio of the coefficient of $(\alpha_s/\pi)^2$ to the one of $\alpha_s/\pi$, the difference is even more pronounced; the corresponding numbers for Eqs.~(\ref{duos}) and (\ref{drhoos}) are roughly 9 versus 5. A phenomenologically interesting application of Eq.~(\ref{lllh}) is to study the effect of QCD corrections on $\Gamma(H\to\ell^+\ell^-)$. The corrections through ${\cal O}(\alpha_s^2G_FM_t^2)$ to this observable may be accommodated by multiplying the Born formula \cite{hff} with \begin{eqnarray} \label{kllh} K_{\ell\ell H}&\hspace*{-2.5mm}=\hspace*{-2.5mm}&(1+\delta_u)^2\nonumber\\ &\hspace*{-2.5mm}=\hspace*{-2.5mm}&1+2\delta_u, \end{eqnarray} where we have suppressed terms of ${\cal O}(G_F^2M_t^4)$ in the second line. This implies that $\delta_u$ is gauge independent and RG invariant in these orders. In order to avoid double counting, the ${\cal O}(G_FM_t^2)$ term must once be subtracted when the full one-loop correction \cite{hff} is included. A detailed numerical analysis will be presented in Section~4. Next, we shall derive the ${\cal O}(\alpha_s^2G_FM_t^2)$ correction to the low-$M_H$ effective $W^+W^-H$ interaction Lagrangian. In contrast to the $\ell^+\ell^-H$ case, we are now faced with the task of computing genuine three-point amplitudes at three loops, which, at first sight, appears to be enormously hard. Fortunately, in the limit that we are interested in, this problem may be reduced to one involving just three-loop two-point diagrams by means of a low-energy theorem \cite{ell,vai}. Generally speaking, this theorem relates the amplitudes of two processes which differ by the insertion of an external Higgs-boson line carrying zero four-momentum. It allows us to compute a loop amplitude, ${\cal M}(A\to B+H)$, with an external Higgs boson which is light compared to the virtual particles by differentiating the respective amplitude without that Higgs boson, ${\cal M}(A\to B)$, with respect to the virtual-particle masses. More precisely \cite{ell,vai}, \begin{equation} \label{let} \lim_{p_H\to0}{\cal M}(A\to B+H)={1\over v}\sum_i {m_i\partial\over\partial m_i}{\cal M}(A\to B), \end{equation} where $i$ runs over all massive virtual particles which are involved in the transition $A\to B$. Here, it is understood that the differential operator does not act on factors of $m_i$ appearing in coupling constants, since this would generate tree-level interactions involving the Higgs boson that do not exist in the SM. This theorem has variously been applied at leading order \cite{ell,vai} and has even made its way into standard text books \cite{oku}. Special care must be exercised if this theorem is to be applied beyond leading order. Then, it must be formulated for the bare quantities of the theory, and the renormalization must be performed after the left-hand side of Eq.~(\ref{let}) has been constructed \cite{ks1}. The beyond-leading-order version of this theorem \cite {ks1} has recently been employed to find the ${\cal O}(\alpha_sG_FM_t^2)$ corrections to $\Gamma\left(H\to b\bar b\,\right)$ \cite{ks1}, $\Gamma\left(Z\to f\bar fH\right)$, and $\sigma(e^+e^-\to ZH)$ \cite{ks2}. A comprehensive review of higher-order applications of this and related low-energy theorems may be found in Ref.~\cite{ks3}. An axiomatic formulation of these soft-Higgs theorems has recently been introduced in Ref.~\cite{kil}. Proceeding along the lines of Refs.~\cite{ks2,ks3}, we find the bare $W^+W^-H$ interaction Lagrangian including its genuine vertex corrections to be \begin{equation} {\cal L}_{W^+W^-H}=2(M_W^0)^2(W_\mu^+)^0(W^{-\mu})^0{H^0\over v^0} \left[1-{(m_t^0)^2\partial\over\partial(m_t^0)^2}\, {\Pi_{WW}(0)\over(M_W^0)^2}\right], \end{equation} where it is understood that $\Pi_{WW}(0)$ is expressed in terms of the bare top-quark mass, $m_t^0$, while all other quark masses are put to zero. We renormalize the $W$-boson mass and wave function according to the electroweak on-shell scheme by substituting \begin{eqnarray} (M_W^0)^2&\hspace*{-2.5mm}=\hspace*{-2.5mm}&M_W^2+\delta M_W^2,\nonumber\\ (W_\mu^\pm)^0&\hspace*{-2.5mm}=\hspace*{-2.5mm}&W_\mu^\pm(1+\delta Z_W)^{1/2}, \end{eqnarray} with the counterterms \begin{eqnarray} \delta M_W^2&\hspace*{-2.5mm}=\hspace*{-2.5mm}&\Pi_{WW}(0),\nonumber\\ \delta Z_W&\hspace*{-2.5mm}=\hspace*{-2.5mm}&-\Pi_{WW}^\prime(0). \end{eqnarray} {}For dimensional reasons, $\delta Z_W$ does not receive corrections in the orders that we are interested in. Using Eq.~(\ref{du}), we thus obtain \begin{equation} \label{lwwh} {\cal L}_{W^+W^-H}=2^{5/4}G_F^{1/2}M_W^2W_\mu^+W^{-\mu}H(1+\delta_{WWH}), \end{equation} where \begin{equation} \delta_{WWH}=\delta_u+\delta_{nu}^{WWH} \end{equation} and the non-universal part herein may be calculated from \begin{equation} \label{dwwhnu} \delta_{nu}^{WWH}=\left[1-{(m_t^0)^2\partial\over\partial(m_t^0)^2}\right] {\Pi_{WW}(0)\over(M_W^0)^2}. \end{equation} In Ref.~\cite{avd}, $\Pi_{WW}(0)$ is expressed in terms of renormalized parameters. Thus, we have to undo the top-quark mass renormalization \cite{gra} before we can apply Eq.~(\ref{dwwhnu}). Then, after evaluating the right-hand side of Eq.~(\ref{dwwhnu}), we reintroduce the renormalized top-quark mass and so obtain a finite result for $\delta_{nu}^{WWH}$, which we combine with $\delta_u$ to get $\delta_{WWH}$. If we define the top-quark mass according to the $\overline{\mbox{MS}}$ scheme, then the result is \begin{eqnarray} \bar\delta_{WWH}&\hspace*{-2.5mm}=\hspace*{-2.5mm}&N_cx_t\left\{-{5\over6} +hC_F\left(-5l-2\zeta(2)+{7\over3}\right) +h^2C_F\left[243S_2-{449\over6}\zeta(3)-{14\over3}\zeta(2)+{79\over3} \right.\right. \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_F\left(-15l^2+l\left(-12\zeta(2)+{83\over2}\right)+4B_4+2D_3 -{1053\over2}S_2+2\zeta(4)+{383\over3}\zeta(3) \right.\nonumber\\ &\hspace*{-2.5mm}-\hspace*{-2.5mm}&\left. {43\over9}\zeta(2) +{377\over72}\right) +C_A\left(-{55\over6}l^2+l\left(-{22\over3}\zeta(2)-{331\over18}\right) -2B_4-D_3+{1053\over4}S_2 \right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left. 15\zeta(4)-{293\over6}\zeta(3)-{29\over3}\zeta(2) -{691\over24}\right) +n_f\left({5\over3}l^2+l\left({4\over3}\zeta(2)+{11\over9}\right) -{8\over3}\zeta(3) \right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left.\left.\left. 2\zeta(2)+{53\over12}\right)\right]\right\} \nonumber\\ &\hspace*{-2.5mm}\approx\hspace*{-2.5mm}& -{5\over6}N_cx_t\left[1+a(2\,l+0.382\,614) +a^2\left({15\over4}l^2+4.184\,802\,l+1.343\,710\right)\right]. \end{eqnarray} The corresponding formula written in terms of $M_t$ reads \begin{eqnarray} \label{dwwhos} \delta_{WWH}&\hspace*{-2.5mm}=\hspace*{-2.5mm}&N_cX_t\left\{-{5\over6} +hC_F\left(-2\zeta(2)+9\right) +h^2C_F\left[243S_2-{449\over6}\zeta(3)+{46\over3}\zeta(2)+{49\over3} \right.\right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_F\left(4B_4+2D_3-{1053\over2}S_2+2\zeta(4)+{443\over3}\zeta(3) +\zeta(2)\left(-80\ln2+{551\over9}\right)-{614\over9}\right) \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_A\left(L\left(-{22\over3}\zeta(2)+33\right) -2B_4-D_3+{1053\over4}S_2+15\zeta(4)-{353\over6}\zeta(3) \right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left.\left.\left. \zeta(2)(40\ln2-23)+{1741\over36}\right) +n_f\left(L\left({4\over3}\zeta(2)-6\right) -{8\over3}\zeta(3)-{14\over3}\zeta(2)-{49\over9}\right)\right]\right\} \nonumber\\ &\hspace*{-2.5mm}\approx\hspace*{-2.5mm}&-{5\over6}N_cX_t[1-2.284\,053\,a-a^2(3.997\,092\,L+10.816\,384)]. \end{eqnarray} We recover the ${\cal O}(G_FM_t^2)$ and ${\cal O}(\alpha_sG_FM_t^2)$ terms of Refs.~\cite{hww,ks3}, respectively. Similarly to $\Delta\rho$ and $\delta_u$, the ${\cal O}(\alpha_s^2G_FM_t^2)$ term of Eq.~(\ref{dwwhos}) supports the screening of the one-loop $M_t$ dependence by the leading-order QCD correction. Here, the coefficient of $(\alpha_s/\pi)^2$ is by 26\% smaller than in the case of $\Delta\rho$, but it, too, is about five times bigger than the coefficient of $\alpha_s/\pi$. {}From Eq.~(\ref{lwwh}) it follows on that $\Gamma(H\to W^+W^-)$ receives the correction factor \begin{equation} \label{kwwh} K_{WWH}=1+2\delta_{WWH}. \end{equation} Thus, both $\delta_{nu}^{WWH}$ and $\delta_{WWH}$ are gauge independent and RG invariant to the orders that we are working in. The tree-level formula for $\Gamma(H\to W^+W^-)$ and its full one-loop correction may be found in Ref.~\cite{hww}. In order for the Higgs boson to decay into a $W^+W^-$ pair, it must satisfy $M_H>2M_W$. On the other hand, the high-$M_t$ approximation is based on $M_H\ll M_t$. Since these two conditions conflict with each other \cite{abe}, the application of Eq.~(\ref{lwwh}) to $\Gamma(H\to W^+W^-)$ is somewhat academic. However, the first condition is relaxed to $M_H>M_W$ or removed altogether if one or both of the $W$ bosons are allowed to leave their mass shells, respectively. In order to avoid gluon exchange between the $W^+W^-H$ vertex and the external fermions, we restrict our considerations to leptonic currents. The resulting class of processes includes $H\to(W^+)^*W^-\to\ell^+\nu_\ell W^-$, $H\to W^+(W^-)^*\to W^+\ell^-\bar\nu_\ell$, $H\to(W^+)^*(W^-)^*\to\ell^+\nu_\ell\ell^{\prime-}\bar\nu_{\ell^\prime}$, as well as $e^+e^-\to\bar\nu_e\nu_e(W^+)^*(W^-)^*\to\bar\nu_e\nu_eH$ via $W^+W^-$ fusion. The Born formulae for these $1\to3$, $1\to4$, and $2\to3$ processes may be found in Refs.~\cite{riz,hzgg,eezh}, respectively. Since $G_F$ is defined through the radiative correction to the muon decay, which is a charged-current process, the $W$-boson propagator does not receive radiative corrections in the orders of interest here. Therefore, the correction factors of all these processes coincide with the one for $\Gamma(H\to W^+W^-)$. Finally, we shall treat the $ZZH$ interaction. The procedure is very similar to the $W^+W^-H$ case. Application of the low-energy theorem (\ref{let}) to the bare $Z$-boson vacuum polarization induced by the top quark yields \begin{equation} {\cal L}_{ZZH}=(M_Z^0)^2Z_\mu^0Z^{\mu0}{H^0\over v^0} \left[1-{(m_t^0)^2\partial\over\partial(m_t^0)^2}\, {\Pi_{ZZ}(0)\over(M_Z^0)^2}\right]. \end{equation} Again, $(M_Z^0)^2=M_Z^2+\delta M_Z^2$, with $\delta M_Z^2=\Pi_{ZZ}(0)$, and $Z_\mu^0=Z_\mu$. Together with Eq.~(\ref{du}), we then have \begin{equation} \label{lzzh} {\cal L}_{ZZH}=2^{1/4}G_F^{1/2}M_Z^2Z_\mu Z^\mu H(1+\delta_{ZZH}), \end{equation} where $\delta_{ZZH}=\delta_u+\delta_{nu}^{ZZH}$, with the non-universal part, \begin{equation} \delta_{nu}^{ZZH}=\left[1-{(m_t^0)^2\partial\over\partial(m_t^0)^2}\right] {\Pi_{ZZ}(0)\over(M_Z^0)^2}, \end{equation} being separately finite, gauge independent, and RG invariant. Starting from the expression for $\Pi_{ZZ}(0)$ listed in Ref.~\cite{avd} and repeating the steps of the $W^+W^-H$ analysis, we obtain \begin{eqnarray} \bar\delta_{ZZH}&\hspace*{-2.5mm}=\hspace*{-2.5mm}&N_cx_t\left\{-{5\over6} +hC_F\left(-5l-2\zeta(2)+{25\over3}\right) +h^2C_F\left[243S_2-{449\over6}\zeta(3)-{14\over3}\zeta(2)+{79\over3} \right.\right. \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_F\left(-15l^2+l\left(-12\zeta(2)+{155\over2}\right)+4B_4+2D_3 -{1053\over2}S_2+2\zeta(4)+{383\over3}\zeta(3) \right.\nonumber\\ &\hspace*{-2.5mm}-\hspace*{-2.5mm}&\left. {259\over9}\zeta(2)+{593\over72}\right) +C_A\left(-{55\over6}l^2+l\left(-{22\over3}\zeta(2)+{65\over18}\right) -2B_4-D_3+{1053\over4}S_2 \right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left. 15\zeta(4)-{293\over6}\zeta(3)-{73\over3}\zeta(2) +{613\over24}\right) \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left.\left. n_f\left({5\over3}l^2+l\left({4\over3}\zeta(2)-{25\over9}\right) -{8\over3}\zeta(3)+{14\over3}\zeta(2)-{35\over12}\right)\right]\right\} \nonumber\\ &\hspace*{-2.5mm}\approx\hspace*{-2.5mm}& -{5\over6}N_cx_t\left[1+a(2\,l-2.017\,386) +a^2\left({15\over4}l^2-4.815\,198\,l-1.086\,685\right)\right] \end{eqnarray} in the $\overline{\mbox{MS}}$ scheme and \begin{eqnarray} \label{dzzhos} \delta_{ZZH}&\hspace*{-2.5mm}=\hspace*{-2.5mm}&N_cX_t\left\{-{5\over6} +hC_F\left(-2\zeta(2)+15\right) +h^2C_F\left[243S_2-{449\over6}\zeta(3)+{46\over3}\zeta(2)+{49\over3} \right.\right. \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_F\left(4B_4+2D_3-{1053\over2}S_2+2\zeta(4)+{443\over3}\zeta(3) +\zeta(2)\left(-80\ln2+{335\over9}\right)-{1019\over9}\right) \nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&C_A\left(L\left(-{22\over3}\zeta(2)+55\right) -2B_4-D_3+{1053\over4}S_2+15\zeta(4)-{353\over6}\zeta(3) \right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left.\left.\left.\! \zeta(2)\left(40\ln2-{113\over3}\right)+{3697\over36}\right) +n_f\left(L\left({4\over3}\zeta(2)-10\right) -{8\over3}\zeta(3)-2\zeta(2)-{115\over9}\right)\right]\right\} \nonumber\\ &\hspace*{-2.5mm}\approx\hspace*{-2.5mm}&-{5\over6}N_cX_t[1-4.684\,053\,a-a^2(8.197\,092\,L+6.846\,779)] \end{eqnarray} in the on-shell scheme. The ${\cal O}(G_FM_t^2)$ and ${\cal O}(\alpha_sG_FM_t^2)$ terms of Eq.~(\ref{dzzhos}) agree with those found in Refs.~\cite{hzz,ks2}, respectively. Again, the ${\cal O}(\alpha_s^2G_FM_t^2)$ term of Eq.~(\ref{dzzhos}) reinforces the potential of the QCD corrections to reduce the leading-order $M_t$ dependence. Comparing $\delta_{ZZH}$ with $\Delta\rho$, $\delta_u$, and $\delta_{WWH}$, we observe that it has the largest $\alpha_s/\pi$ coefficient but the smallest $(\alpha_s/\pi)^2$ coefficient, the ratio of the latter to the former only being about 1.5. The $(\alpha_s/\pi)^2$ coefficient of $\delta_{ZZH}$ is by 53\% smaller than the one of $\Delta\rho$ in Eq.~(\ref{drhoos}). {}From Eq.~(\ref{lzzh}), we infer that $\Gamma(H\to ZZ)$ receives the correction factor \begin{equation} \label{kzzh} K_{ZZH}=1+2\delta_{ZZH}. \end{equation} The Born formula for $\Gamma(H\to ZZ)$ and its full one-loop correction may be found in Ref.~\cite{hzz}. Since the condition $2M_Z<M_H\ll M_t$ is likely to be unrealistic \cite{abe}, the high-$M_t$ approximation underlying Eq.~(\ref{lzzh}) is of limited usefulness for $H\to ZZ$. We can evade this problem by allowing for one or both of the $Z$ bosons to go off-shell. In addition to the information contained in Eq.~(\ref{lzzh}), we then need to account for the corresponding corrections arising from the gauge sector. However, in order not to invoke unknown QCD corrections, we have to restrict ourselves to the inclusion of lepton lines. The form of the additional corrections depends on the considered reaction. It is useful to divide the phenomenologically relevant processes into three classes: \begin{itemize} \item[(1)] $H\to Z^*Z\to f\bar fZ$, $Z\to Z^*H\to f\bar fH$, and $e^+e^-\to ZH$; \item[(2)] $H\to Z^*Z^*\to f\bar ff^\prime\bar f^\prime$ and $e^+e^-\to Z^*\to Z^*H\to f\bar fH$ (via Higgs-strahlung); \item[(3)] $e^+e^-\to e^+e^-Z^*Z^*\to e^+e^-H$ (via $ZZ$ fusion). \end{itemize} Here, $f$ and $f^\prime$ stand for neutrinos and charged leptons. The results for $H\to f\bar fZ$ at tree level \cite{riz} and at one loop \cite{pr}, for $Z\to f\bar fH$ at tree level \cite{zhgg} and at one loop \cite{zffh}, for $e^+e^-\to ZH$ at tree level and at one loop \cite{eezh}, for $H\to f\bar ff^\prime\bar f^\prime$ at tree level \cite{hzgg}, and for $e^+e^-\to f\bar fH$ at tree level \cite{eezh} are in the literature. In the next section, we shall discuss the corrections to these processes in ${\cal O}(\alpha_s^nG_FM_t^2)$, with $n=0,1,2$. \section{Corrections from the gauge sector} The IBA \cite{iba} provides a systematic and convenient method to incorporate the dominant corrections of fermionic origin to processes within the gauge sector of the SM. These are contained in $\Delta\rho$ and $\Delta\alpha=1-\alpha/\overline\alpha$, which parameterizes the running of the fine-structure constant from its value, $\alpha$, defined in Thomson scattering to its value, $\overline\alpha$, measured at the $Z$-boson scale. The recipe is as follows. Starting from the Born formula expressed in terms of $\alpha$, $c_w$, $s_w$, and the physical particle masses, one substitutes \begin{eqnarray} \alpha&\hspace*{-2.5mm}\to\hspace*{-2.5mm}&\overline\alpha={\alpha\over1-\Delta\alpha},\nonumber\\ c_w^2&\hspace*{-2.5mm}\to\hspace*{-2.5mm}&\overline c_w^2=1-\overline s_w^2=c_w^2(1-\Delta\rho). \end{eqnarray} To eliminate $\overline\alpha$ in favour of $G_F$, one exploits the relation \begin{equation} {\sqrt2\over\pi}G_F={\overline\alpha\over\overline s_w^2M_W^2} ={\overline\alpha\over\overline c_w^2\overline s_w^2M_Z^2}(1-\Delta\rho), \end{equation} which correctly accounts for the leading fermionic corrections. We shall now employ the IBA to find the additional corrections through ${\cal O}(\alpha_s^2G_FM_t^2)$ to the four- and five-point processes with a $ZZH$ coupling, which we have classified in Section~2. We shall always assume that the Born formulae are written in terms of $G_F$, $c_w$, $s_w$, and the physical particle masses. The generic correction factor for class~(1) reads \cite{ks2,zffh} \begin{eqnarray} \label{k1} K_1^{(f)}&\hspace*{-2.5mm}=\hspace*{-2.5mm}&{(1+\delta^{ZZH})^2\over1-\Delta\rho}\, {\overline v_f^2+a_f^2\over v_f^2+a_f^2}\nonumber\\ &\hspace*{-2.5mm}=\hspace*{-2.5mm}&1+2\delta^{ZZH}+ \left(1-8c_w^2{Q_fv_f\over v_f^2+a_f^2}\right)\Delta\rho, \end{eqnarray} where $v_f=2I_f-4s_w^2Q_f$, $\overline v_f=2I_f-4\overline s_w^2Q_f$, $a_f=2I_f$, $Q_f$ is the electric charge of $f$ in units of the positron charge, $I_f$ is the third component of weak isospin of the left-handed component of $f$, and we have omitted terms of ${\cal O}(G_F^2M_t^4)$ in the second line. Similarly, the correction factor for class~(2) is given by \cite{gro} \begin{eqnarray} \label{k2} K_2^{(ff^\prime)}&\hspace*{-2.5mm}=\hspace*{-2.5mm}& {(1+\delta^{ZZH})^2\over(1-\Delta\rho)^2}\, {\overline v_f^2+a_f^2\over v_f^2+a_f^2}\, {\overline v_{f^\prime}^2+a_{f^\prime}^2\over v_{f^\prime}^2+a_{f^\prime}^2} \nonumber\\ &\hspace*{-2.5mm}=\hspace*{-2.5mm}&1+2\delta^{ZZH}+ 2\left[1-4c_w^2\left({Q_fv_f\over v_f^2+a_f^2} +{Q_{f^\prime}v_{f^\prime}\over v_{f^\prime}^2+a_{f^\prime}^2}\right)\right] \Delta\rho. \end{eqnarray} Here and in the following, we neglect interference terms of five-point amplitudes with a single fermion trace, since, in the kinematic regime of interest here, these are strongly suppressed, by $\Gamma_V/M_V$, with $V=W,Z$. Such terms have recently been included in a tree-level calculation of $\Gamma(H\to2V\to4f)$ for $M_H\ll M_W$ \cite{asa}. The correction factor for case~(3) is slightly more complicated because the electron and positron lines run from the initial state to the final state. Allowing for generic fermion flavours, $f$ and $f^\prime$, the Born cross section may be evaluated from \begin{equation} \label{zfus} \sigma(ff^\prime\to ff^\prime H)={G_F^3M_Z^4\over64\pi^3\sqrt2} \left[(v_f^2+a_f^2)(v_{f^\prime}^2+a_{f^\prime}^2)A \pm4v_fa_fv_{f^\prime}a_{f^\prime}B\right], \end{equation} where \begin{equation} \label{int} A=\int_{M_H^2/s}^1dx\int_x^1dy{a(x,y)\over\left[1+s(y-x)/M_Z^2\right]^2}, \end{equation} and similarly for $B$, $\sqrt s$ is the centre-of-mass energy, and the plus/minus sign refers to an odd/even number of antifermions in the initial state. The process under case~(3), with an $e^+e^-$ initial state, requires the plus sign. The integrands read \begin{eqnarray} a(x,y)&\hspace*{-2.5mm}=\hspace*{-2.5mm}& \left({2x\over y^3}-{1+2x\over y^2}+{2+x\over2y}-{1\over2}\right) \left[{z\over1+z}-\ln(1+z)\right] +{x\over y^2}\left({1\over y}-1\right){z^2\over1+z},\nonumber\\ b(x,y)&\hspace*{-2.5mm}=\hspace*{-2.5mm}& \left(-{x\over y^2}+{2+x\over2y}-{1\over2}\right) \left[{z\over1+z}-\ln(1+z)\right], \end{eqnarray} where $z=(y/M_Z^2)(s-M_H^2/x)$. The inner integration in Eq.~(\ref{int}) has been carried out analytically in Appendix~A of Ref.~\cite{eezh}. By means of the IBA, we obtain the correction factor for Eq.~(\ref{zfus}) as \begin{eqnarray} \label{k3} K_3^{(ff^\prime)}&\hspace*{-2.5mm}=\hspace*{-2.5mm}& {(1+\delta^{ZZH})^2\over(1-\Delta\rho)^2}\, {(\overline v_f^2+a_f^2)(\overline v_{f^\prime}^2+a_{f^\prime}^2)A \pm4\overline v_fa_f\overline v_{f^\prime}a_{f^\prime}B\over (v_f^2+a_f^2)(v_{f^\prime}^2+a_{f^\prime}^2)A \pm4v_fa_fv_{f^\prime}a_{f^\prime}B}\nonumber\\ &\hspace*{-2.5mm}=\hspace*{-2.5mm}&1+2\delta^{ZZH}+ 2\left[1-{4c_w^2\over1+r}\left({Q_fv_f\over v_f^2+a_f^2} +{Q_{f^\prime}v_{f^\prime}\over v_{f^\prime}^2+a_{f^\prime}^2}\right) \right.\nonumber\\&\hspace*{-2.5mm}-\hspace*{-2.5mm}&\left. {2c_w^2\over1+1/r}\left({Q_f\over v_f}+{Q_{f^\prime}\over v_{f^\prime}}\right) \right]\Delta\rho, \end{eqnarray} where \begin{equation} r={\pm4v_fa_fv_{f^\prime}a_{f^\prime}B\over (v_f^2+a_f^2)(v_{f^\prime}^2+a_{f^\prime}^2)A}. \end{equation} We wish to point out that, in the limit $r\to0$, $K_3$ coincides with $K_2$. Detailed analysis reveals that $r$ is quite small in magnitude whenever $e^+e^-\to e^+e^-H$ via $ZZ$ fusion is phenomenologically relevant. In fact, if we consider energies $\sqrt s>150$~GeV and demand that the total cross section of this process be in excess of $10^{-2}$~fb$^{-1}$, then we find $|r|<1\%$. This concludes our discussion of the additional QCD corrections to the processes under items~(1)--(3) originating in the gauge sector. \section{Numerical results} We are now in a position to explore the phenomenological implications of our results. We shall take the values of our input parameters to be $M_W=80.26$~GeV, $M_Z=91.1887$~GeV \cite{lep}, $M_t=180$~GeV \cite{abe}, and $\alpha_s^{(5)}(M_Z)=0.118$ \cite{bet}.\footnote{Note that this value does not include results from lattice computations.} The latter corresponds to $\alpha_s^{(6)}(M_t)=0.1071$, which entails that $\Lambda_{\overline{MS}}^{(6)}=91$~MeV in Eq.~(\ref{as}). If we use the one-loop formula for $\alpha_s^{(6)}(\mu)$, {\it i.e.}, Eq.~(\ref{as}) with the second term within the square brackets discarded, $\Lambda_{\overline{MS}}^{(6)}$ comes down to 41~MeV. Any perturbative calculation to finite order depends on the choice of renormalization scheme and, in general, also on one or more renormalization scales. It is generally believed that the scheme and scale dependences of a calculation up to a given order indicate the size of the unknown higher-order contributions, {\it i.e.}, they provide us with an estimate of the theoretical uncertainty. Of course, the central values and variations of the scales must be judiciously chosen in order for this estimate to be meaningful. If the perturbation series converges, then the scheme and scale dependences are expected to decrease as the respective next order is taken into account. This principle has recently been confirmed for $\Delta\rho$ \cite{avd,yr}. Here, we have the opportunity to carry out similar studies for the three additional observables $\delta_u$, $\delta_{WWH}$, and $\delta_{ZZH}$. Similarly to Ref.~\cite{avd}, we have presented our results in the on-shell and $\overline{\mbox{MS}}$ schemes as functions of a single renormalization scale, $\mu$. In the $\overline{\mbox{MS}}$ scheme, one could, in principle, introduce individual renormalization scales for the coupling and the mass. {}For simplicity, we have chosen not to do so. It is natural to define the central value of $\mu$ in such a way that, at this point, the radiative correction is devoid of logarithmic terms. This leads us to set $\mu=\xi M_t$ in the on-shell scheme and $\mu=\xi\mu_t$, where $\mu_t=m_t(\mu_t)$, in the $\overline{\mbox{MS}}$ scheme. We may obtain $\mu_t$ as a closed function of $M_t$ by iterating Eq.~(\ref{mass}), with the result that \begin{eqnarray} \label{mut} {\mu_t\over M_t}&\hspace*{-2.5mm}=\hspace*{-2.5mm}&1-4HC_F +H^2C_F\left\{-12\zeta(2)+6 +C_F\left[-12\zeta(3)+6\zeta(2)(8\ln2-5)+{199\over8}\right] \right.\nonumber\\ &\hspace*{-2.5mm}+\hspace*{-2.5mm}&\left. C_A\left[6\zeta(3)+8\zeta(2)(-3\ln2+1)-{1111\over24}\right] +n_f\left[4\zeta(2)+{71\over12}\right]\right\} \nonumber\\ &\hspace*{-2.5mm}\approx\hspace*{-2.5mm}&1-{4\over3}A-6.458\,784\,A^2, \end{eqnarray} where $A=4H=\alpha_s(M_t)/\pi$. {}For $M_t=180$~GeV, Eq.~(\ref{mut}) yields $\mu_t=170.5$~GeV, in good agreement with the exact fix point of Eq.~(\ref{mass}), which is $\mu_t=170.6$~GeV. \begin{table}[ht] \caption{Relative deviations (in \%) of $\Delta\rho$, $\delta_u$, $\delta_{WWH}$, and $\delta_{ZZH}$ from the respective one-loop results due to their corrections up to ${\cal O}(\alpha_s)$ and ${\cal O}(\alpha_s^2)$. The renormalization scale dependence is investigated by choosing $\mu=\xi M_t$, with $\xi$ variable. }\label{tab:os} \medskip \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline\hline \rule{0mm}{5mm}$\xi$ & \multicolumn{2}{c|}{$\Delta\rho/\Delta\rho^{(1)}-1$ [\%]} & \multicolumn{2}{c|}{$\delta_u/\delta_u^{(1)}-1$ [\%]} & \multicolumn{2}{c|}{$\delta_{WWH}/\delta_{WWH}^{(1)}-1$ [\%]} & \multicolumn{2}{c|}{$\delta_{ZZH}/\delta_{ZZH}^{(1)}-1$ [\%]} \\ \cline{2-9} \rule{0mm}{5mm} & ${\cal O}(\alpha_s)$ & ${\cal O}(\alpha_s^2)$ & ${\cal O}(\alpha_s)$ & ${\cal O}(\alpha_s^2)$ & ${\cal O}(\alpha_s)$ & ${\cal O}(\alpha_s^2)$ & ${\cal O}(\alpha_s)$ & ${\cal O}(\alpha_s^2)$ \\ \hline 1/4 & $-11.68$ & $-11.88$ & $ -7.34$ & $ -8.65$ & $ -9.33$ & $ -9.35$ & $-19.13$ & $-16.58$ \\ 1/2 & $-10.63$ & $-11.72$ & $ -6.68$ & $ -8.34$ & $ -8.49$ & $ -9.24$ & $-17.40$ & $-16.83$ \\ 1 & $ -9.75$ & $-11.44$ & $ -6.12$ & $ -8.01$ & $ -7.78$ & $ -9.04$ & $-15.96$ & $-16.76$ \\ 2 & $ -9.00$ & $-11.11$ & $ -5.66$ & $ -7.67$ & $ -7.19$ & $ -8.79$ & $-14.74$ & $-16.51$ \\ 4 & $ -8.36$ & $-10.74$ & $ -5.26$ & $ -7.35$ & $ -6.68$ & $ -8.51$ & $-13.70$ & $-16.15$ \\ \hline\hline \end{tabular} \end{table} \begin{table}[ht] \caption{Relative deviations (in \%) of $\Delta\bar\rho$, $\bar\delta_u$, $\bar\delta_{WWH}$, and $\bar\delta_{ZZH}$ from the respective one-loop results due to their corrections up to ${\cal O}(\alpha_s)$ and ${\cal O}(\alpha_s^2)$. The renormalization scale dependence is investigated by choosing $\mu=\xi\mu_t$, with $\xi$ variable. }\label{tab:ms} \medskip \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline\hline \rule{0mm}{5mm}$\xi$ & \multicolumn{2}{c|}{$\Delta\bar\rho/\Delta\rho^{(1)}-1$ [\%]} & \multicolumn{2}{c|}{$\bar\delta_u/\delta_u^{(1)}-1$ [\%]} & \multicolumn{2}{c|}{$\bar\delta_{WWH}/\delta_{WWH}^{(1)}-1$ [\%]} & \multicolumn{2}{c|}{$\bar\delta_{ZZH}/\delta_{ZZH}^{(1)}-1$ [\%]} \\ \cline{2-9} \rule{0mm}{5mm} & ${\cal O}(\alpha_s)$ & ${\cal O}(\alpha_s^2)$ & ${\cal O}(\alpha_s)$ & ${\cal O}(\alpha_s^2)$ & ${\cal O}(\alpha_s)$ & ${\cal O}(\alpha_s^2)$ & ${\cal O}(\alpha_s)$ & ${\cal O}(\alpha_s^2)$ \\ \hline 1/4 & $-15.63$ & $-11.15$ & $-10.70$ & $ -8.25$ & $-12.96$ & $ -8.67$ & $-24.09$ & $-15.15$ \\ 1/2 & $-10.89$ & $-11.55$ & $ -6.87$ & $ -8.22$ & $ -8.71$ & $ -9.09$ & $-17.77$ & $-16.56$ \\ 1 & $ -8.96$ & $-11.19$ & $ -5.63$ & $ -7.79$ & $ -7.15$ & $ -8.86$ & $-14.69$ & $-16.51$ \\ 2 & $ -8.64$ & $-10.88$ & $ -5.82$ & $ -7.56$ & $ -7.11$ & $ -8.71$ & $-13.49$ & $-16.16$ \\ 4 & $ -9.24$ & $-10.85$ & $ -6.81$ & $ -7.70$ & $ -7.92$ & $ -8.84$ & $-13.41$ & $-15.93$ \\ \hline\hline \end{tabular} \end{table} In Tables~\ref{tab:os} and \ref{tab:ms}, we investigate the $\xi$ dependence of $\delta_u$, $\delta_{WWH}$, and $\delta_{ZZH}$ and their $\overline{\mbox{MS}}$ counterparts, respectively. {}For comparison, we also include the results for $\Delta\rho$ and $\Delta\bar\rho$. To be specific, we consistently evaluate these quantities to leading and next-to-leading order in QCD and study their relative deviations from their respective one-loop values, which we denote by the superscript (1), {\it e.g.}, $\Delta\rho^{(1)}=N_cX_t$, {\it etc.} Notice that the on-shell and $\overline{\mbox{MS}}$ results coincide at one loop. In our ${\cal O}(\alpha_s)$ analysis, we use the one-loop formula for $\alpha_s(\mu)$ with $\Lambda_{\overline{MS}}^{(6)}=41$~MeV and omit the ${\cal O}(\alpha_s^2)$ terms in Eqs.~(\ref{mass}) and (\ref{mut}). {}For the time being, let us concentrate on the entries for $\xi=1$ and assess the effect of the QCD corrections as well as their scheme dependence. We observe that, in both schemes, the QCD corrections are throughout negative, even for $\bar\delta_u$ and $\bar\delta_{WWH}$, where the ${\cal O}(\alpha_s)$ and ${\cal O}(\alpha_s^2)$ terms are in part positive. This is due to the fact that we consistently compute all QCD parameters, {\it i.e.}, $\alpha_s(\mu)$, $m_t(\mu)$, and $\mu_t$, to the orders under consideration. The reduction in $x_t$, which occurs as an overall factor in the $\overline{\mbox{MS}}$ formulae, happens to overcompensate the positive effect of these particular coefficients. Inclusion of the ${\cal O}(\alpha_s^2)$ terms in $(\Delta\rho,\delta_u,\delta_{WWH},\delta_{ZZH})$ increases the size of the QCD corrections by $(17,31,16,5)\%$, respectively. In the $\overline{\mbox{MS}}$ case, the increments amount to $(25,38,24,12)\%$ of the respective ${\cal O}(\alpha_s)$ corrections. As might be expected, the scheme dependence of the QCD corrections to this quadruplet of quantities is dramatically reduced, by $(68,55,71,80)\%$, as we pass from ${\cal O}(\alpha_s)$ to ${\cal O}(\alpha_s^2)$. Let us now also include the other $\xi$ values in our consideration. Within each scheme, we determine the scale dependence of the QCD correction to a given quantity by comparing its largest and smallest values in the interval $1/4\le\xi\le4$. As expected, the scale dependence is drastically decreased when we take the ${\cal O}(\alpha_s^2)$ terms into account, namely by $(66,37,68,87)\%$ and $(89,87,93,86)\%$ in the on-shell and $\overline{\mbox{MS}}$ schemes, respectively. The exceptionally small reduction of the scale dependence in the case of $\delta_u$ is due to the fact that $\delta_u$ has the smallest ${\cal O}(\alpha_s)$ term and the largest ${\cal O}(\alpha_s^2)$ term of all four on-shell quantities. \begin{table}[ht] \caption{Coefficients of the correction factors in the form of Eq.~(\ref{kfac}) for the various Higgs-boson decay rates and production cross sections discussed in the text. In the last line, $x=B/A$, where $A$ and $B$ are given by Eq.~(\ref{int}), and terms of ${\cal O}(x^2)$ have been neglected. }\label{tab:k} \medskip \begin{tabular}{|c|c|c|c|} \hline\hline $K$ & $C_1$ & $C_2$ & $C_3$ \\ \hline $K_{\ell\ell H}$ & $7/3$ & $-1.797$ & $-16.201$ \\ $K_{WWH}$ & $-5/3$ & $-2.284$ & $-10.816$ \\ $K_{ZZH}$ & $-5/3$ & $-4.684$ & $-6.847$ \\ $K_1^{(\nu)}$ & $-2/3$ & $-7.420$ & $4.774$ \\ $K_1^{(\ell)}$ & $-1.272$ & $-5.249$ & $-4.445$ \\ $K_2^{(\nu\nu)}$ & $1/3$ & $6.261$ & $-53.330$ \\ $K_2^{(\nu\ell)}$ & $-0.272$ & $-14.025$ & $32.824$ \\ $K_2^{(\ell\ell)}$ & $-0.878$ & $-6.323$ & $0.113$ \\ $K_3^{(\ell\ell)}$ & $-0.878-2.353\,x$ & $-6.323+9.281\,x$ & $0.113-39.416\,x$ \\ \hline\hline \end{tabular} \end{table} In the remainder of this section, we shall stick to the on-shell scheme. In Eqs.~(\ref{kllh}), (\ref{kwwh}), (\ref{kzzh}), (\ref{k1}), (\ref{k2}), and (\ref{k3}), we have presented correction factors for various Higgs-boson production cross sections and decay rates in terms of $\Delta\rho$, $\delta_u$, $\delta_{WWH}$, and $\delta_{ZZH}$. It is instructive to cast these correction factors into the generic form \begin{equation} \label{kfac} K=1+C_1\Delta\rho^{(1)}\left[1+C_2a\left(1+{7\over4}aL\right)+C_3a^2\right], \end{equation} where $C_i$ $(i=1,2,3)$ are numerical coefficients. Notice that we have kept the full $\mu$ dependence in Eq.~(\ref{kfac}). We could have written Eqs.~(\ref{drhoos}), (\ref{duos}), (\ref{dwwhos}), and (\ref{dzzhos}) in the same way. The fact that the coefficient of $aL$ is universal may be understood by observing that $K$ represents a physical observable, which must be RG invariant through the order of our calculation, and that, to leading order of QCD, $K$ only implicitly depends on $\mu$, via $a$. In fact, the coefficient of $aL$ is nothing but $\beta_0$ of Eq.~(\ref{beta}). The outcome of this decomposition is displayed in Table~\ref{tab:k}. In the case of $K_3^{(\ell\ell)}$, we have treated $x=B/A$, where $A$ and $B$ are defined in Eq.~(\ref{int}), as an additional expansion parameter and discarded terms of ${\cal O}(x^2)$. This is justified because, in practice, $|x|\ll1$, {\it e.g.}, for $\sqrt s=300$~GeV and $M_H=100$~GeV, we find $x\approx-5.233\cdot10^{-2}$. While in the case of the three basic corrections, $K_{\ell\ell H}$, $K_{WWH}$, and $K_{ZZH}$, $C_2$ and $C_3$ are both negative, this is not in general so. In fact, in all composite corrections, except for $K_1^{(\ell)}$, the ${\cal O}(\alpha_s^2)$ terms partially compensate the ${\cal O}(\alpha_s)$ ones. In $K_2^{(\nu\nu)}$, we even find a counterexample to the heuristic rule \cite{ks2} that, in the $G_F$ formulation of the on-shell scheme, the ${\cal O}(G_FM_t^2)$ terms get screened by their QCD corrections. In the latter case, we also encounter a gigantic value of $C_3$. Both features may be ascribed to the fact, that, in ${\cal O}(G_FM_t^2)$, the $\delta_{ZZH}$ and $\Delta\rho$ terms of $K_2^{(\nu\nu)}$ largely cancel. The extraordinarily large value of $C_3$ in $K_2^{(\nu\ell)}$ is also accompanied by a suppression of $C_1$. The $C_1$ values of $K_2^{(\ell\ell)}$ and $K_3^{(\ell\ell)}$ are relatively small, too. We are thus in the fortunate position that the leading high-$M_t$ corrections to the $2\to3$ and $1\to4$ processes of Higgs-boson production and decay with a $ZZH$ coupling, for which full one-loop calculations have not yet been performed, are throughout quite small. Thus, there is hope that the subleading fermionic corrections to these processes will not drastically impair the situation. However, the IBA does not provide us with any information on the bosonic corrections. \begin{table}[ht] \caption{Full one-loop weak corrections (in \%) to various Higgs-boson decay rates and production cross sections and their ${\cal O}(G_FM_t^2)$ terms. In the last line, we have used $\protect\sqrt s=175$~GeV. }\label{tab:full} \medskip \begin{tabular}{|c|c|c|c|} \hline\hline Observable & $M_H$ [GeV] & ${\cal O}(\alpha)$ weak [\%] & ${\cal O}(G_FM_t^2)$ [\%] \\ \hline $\Gamma(H\to\tau^+\tau^-)$ & 75 & 1.792 & 2.369 \\ $\Gamma(H\to\nu\bar\nu Z)$ & 105 & 1.275 & $-0.677$ \\ $\Gamma(H\to\ell^+\ell^-Z)$ & 105 & $-1.220$ & $-1.292$ \\ $\Gamma(Z\to\nu\bar\nu H)$ & 65 & 0.024 & $-0.677$ \\ $\Gamma(Z\to\ell^+\ell^-H)$ & 65 & 0.296 & $-1.292$ \\ $\sigma(e^+e^-\to ZH)$ & 75 & $-2.293$ & $-1.292$ \\ \hline\hline \end{tabular} \end{table} In this context, it is interesting to revisit processes for which the full one-loop weak corrections are known and to investigate in how far the ${\cal O}(G_FM_t^2)$ terms play a dominant r\^ole there. Here, we are only interested in reactions which already proceed at tree level. Specifically, we shall consider $Z\to f\bar fH$ \cite{zffh} for $M_H=65$~GeV, $H\to\tau^+\tau^-$ \cite{hff} and $e^+e^-\to ZH$ \cite{eezh} for $M_H=75$~GeV, and $H\to f\bar fZ$ \cite{pr} for $M_H=105$~GeV, where $f=\nu,\ell$. Our analysis of $\sigma(e^+e^-\to ZH)$ will refer to LEP2 energy, $\sqrt s=175$~GeV. In all these cases, the quantumelectrodynamical (QED) and weak corrections are separately finite and gauge independent at one loop. In Table~\ref{tab:full}, we compare the full one-loop weak corrections to these processes with their ${\cal O}(G_FM_t^2)$ terms. In the case of $H\to\tau^+\tau^-$, $H\to\ell^+\ell^-Z$, and $e^+e^-\to ZH$, the ${\cal O}(G_FM_t^2)$ terms give a reasonably good account of the full corrections, while they come out with the wrong sign in the other cases. However, the full calculations for $M_H=65$~GeV give very small results anyway. On the other hand, the ${\cal O}(G_FM_t^2)$ term for $H\to\nu\bar\nu Z$ is suppressed due to a partial cancellation between $\delta_{ZZH}$ and $\Delta\rho$ in $K_1^{(\nu)}$ and cannot be expected to dominate the full correction. Whenever the full correction is known, it should be included on the right-hand side of Eq.~(\ref{kfac}) with the ${\cal O}(G_FM_t^2)$ term subtracted. In conclusion, the radiative corrections considered in Table~\ref{tab:full} all appear to be well under control. \section{Conclusions} In this paper, we have presented the three-loop ${\cal O}(\alpha_s^2G_FM_t^2)$ corrections to the effective Lagrangians for the interactions of light Higgs bosons with pairs of charged leptons, $W$ bosons, and $Z$ bosons in the SM. While the demand for corrections in this order is certainly more urgent in the gauge sector \cite{avd}, where precision test are presently being carried out, our analysis is also interesting from a theoretical point of view, since it allows us to recognize a universal pattern. In addition to $\Delta\rho$, we have now three more independent observables with quadratic $M_t$ dependence at our disposal for which the QCD expansion is known up to next-to-leading order, namely $\delta_u$, $\delta_{WWH}$, and $\delta_{ZZH}$. In the on-shell scheme of electroweak and QCD renormalization, these four electroweak parameters exhibit striking common properties. In fact, the leading- and next-to-leading-order QCD corrections act in the same direction and screen the ${\cal O}(G_FM_t^2)$ terms. Even the sets of $\alpha_s/\pi$ and $(\alpha_s/\pi)^2$ coefficients each lie in the same ball park. {}For the choice $\mu=M_t$, the coefficients of $\alpha_s/\pi$ range between $-1.797$ and $-4.684$, and those of $(\alpha_s/\pi)^2$ between $-6.847$ and $-16.201$. If we compare this with the corresponding coefficients of the ratio $\mu_t^2/M_t^2$, which are $-2.667$ and $-11.140$, then it becomes apparent that the use of the top-quark pole mass is the origin of these similarities. Here, $\mu_t=m_t(\mu_t)$, for which we have presented a closed two-loop formula. If we express the QCD expansions in terms of $\mu_t$ rather than $M_t$ and choose $\mu=\mu_t$, then the coefficients of $\alpha_s/\pi$ and $(\alpha_s/\pi)^2$ nicely group themselves around zero; they range from $-2.017$ to 0.870 and from $-3.970$ to $1.344$, respectively. This indicates that the perturbation expansions converge more rapidly if we renormalize the top-quark mass according to the $\overline{\mbox{MS}}$ scheme. Without going into details, we would like to mention that the study of renormalons \cite{ren} offers a possible theoretical explanation of this observation. Since the on-shell and $\overline{\mbox{MS}}$ results coincide in lowest order, this does, of course, not imply that the QCD corrections are any smaller in the $\overline{\mbox{MS}}$ scheme. It just means that, as a rule, the ${\cal O}(G_FM_t^2)$ terms with $M_t$ replaced by the two-loop expression for $\mu_t$ are likely to provide fair approximations for the full three-loop results. Furthermore, we have demonstrated that, similarly to $\Delta\rho$, the scheme and scale dependences of $\delta_u$, $\delta_{WWH}$, and $\delta_{ZZH}$ are considerably reduced when the next-to-leading-order QCD corrections are taken into account. Armed with this information, we have made rather precise predictions for a variety of production and decay processes of low-mass Higgs bosons at present and future $e^+e^-$ colliders. In all the cases considered here, the radiative corrections appear to be well under control now. \bigskip \centerline{\bf ACKNOWLEDGEMENTS} \smallskip\noindent We would like to thank Bill Bardeen, Kostja Chetyrkin, and Michael Spira for very useful discussions. One of us (BAK) is indebted to the FNAL Theory Group for inviting him as a Guest Scientist. He is also grateful to the Phenomenology Department of the University of Wisconsin at Madison for the great hospitality extended to him during a visit when a major part of his work on this project was carried out.
\section{Introduction} Several aspects of the collective transport associated with the phason in charge- or spin-density waves (CDW or SDW) are still not well understood. One of the intriguing phenomena is the electromechanical effect observed in CDW\cite{Brill,Moz,Xi1,Xi2,Jac} and more recently in SDW\cite{Brown}. First of all, most of the elastic moduli increase upon entrance into the CDW or SDW state, often with a sharp dip at $T_c$, the transition temperature. Second, some of the elastic moduli in CDW or SDW soften when the density wave is depinned by an external electric field in excess of the depinning threshold field $E_T$. Third, the change in the elastic moduli due to depinning of the density wave depends on the frequency $\omega$ of the flexural vibration\cite{Xi3}, and decreases like $\omega^{-p}$ with $p\approx 1$. This behavior is similar to the frequency dependence of the change in the dielectric constant upon depinning in CDW and SDW\cite{Cava}. We have shown earlier\cite{MV1} in the collisionless limit that the hardening of the elastic constants can be understood in terms of the reduction in the quasiparticle screening of the ion potential due to the formation of the density wave state. The electromechanical effect was interpreted as an additional screening contribution from the collective mode of the density wave condensate (phason) liberated by depinning. A later extension of that theory to the experimentally more relevant hydrodynamic limit\cite{VM1} did not modify the above picture qualitatively. However in these papers it was assumed that the phonon simply couples to the electronic density, and the effect of the long-range Coulomb interaction was neglected. The purpose of the present paper is twofold. First, we shall develop the theory of the electron-phonon coupling for a strongly anisotropic system, which will enable us to distinguish between the behavior of transverse and longitudinal sound waves propagating in various directions. Second, we shall include the effect of the long-range Coulomb interaction following Kadanoff and Falko\cite{KF}. In Section II we construct the electronic stress tensor (which couples to the deformation tensor of the sound wave) for a quasi-one dimensional system following the method of comoving frame\cite{Tsu}. In Section III we concentrate on the quasiparticle contribution to the stress tensor correlation functions corresponding to the pinned case. Section IV is devoted to the examination of the coupling of the stress tensor to the phason, which is relevant to the electromechanical effect. Our conclusions are summarized in Section V. A preliminary report on this work has already been published elsewhere\cite{VM2}. \section{Electron-phonon coupling} Following Tsuneto\cite{Tsu} let us assume that both the ionic potential and the electronic wavefunction are deformed in the slowly varying sound field ${\bf u}({\bf r},t)={\bf u}\cos({\bf qr}-\omega t)$ imposed externally (extreme tight-binding limit): \begin{eqnarray} V({\bf r})\rightarrow&V[{\bf r}-{\bf u}({\bf r},t)]\nonumber\\ \psi({\bf r})\rightarrow&\psi[{\bf r}-{\bf u}({\bf r},t)] (1+\nabla{\bf u})^{-1/2}. \end{eqnarray} This displacement is generated by the unitary operator $U=\exp[-{\bf u}({\bf r},t)\nabla]$. Therefore the deformed wavefunction is an eigenfunction of the transformed Hamiltonian $h=Uh_0U^+$, where $h_0=\varepsilon(-i\nabla)$ with $\varepsilon ({\bf p})$ being the zone-periodic electronic energy spectrum. The transformed Hamiltonian $h$ is then expanded in terms of the deformation tensor $\nabla_iu_j$, which is much smaller than one, even though the displacement ${\bf u}$ itself may be many times of the lattice constant for sound propagation. We obtain $h=h_0+h_{el-ph}$, where the Hamiltonian for the electron-phonon coupling is given by \begin{equation} h_{el-ph}=\sum_{i,j}(\nabla_iu_j)i\nabla_jv_i(-i\nabla), \end{equation} with ${\bf v}({\bf p})=\partial\varepsilon({\bf p})/\partial {\bf p}$, the velocity of the Bloch electron. The matrix element of $h_{el-ph}$ between Bloch states is evaluated as \begin{equation} \langle {\bf p}+{\bf q}|h_{el-ph}|{\bf p}\rangle = -i\sum_{i,j}q_ju_i\tau_{ij}({\bf p}), \end{equation} where the stress tensor \begin{equation} \tau_{ij}({\bf p})=mv_i({\bf p})v_j({\bf p}), \end{equation} and $m$ is the bare electron mass. This expression generalizes the stress tensor used for an isotropic metal\cite{KF}. In orthorombic symmetry the sound wave polarized in the ${\bf i}$ direction and propagating in the ${\bf j}$ direction couples to the $\tau_{ij}$ component of the stress tensor, and in order to determine the effect of that coupling on the frequency (or velocity) of the sound, we have to evaluate the appropriate stress tensor correlation function $\langle [\tau_{ij},\tau_{ij}]\rangle$. Once this is known, the renormalized sound velocity can be calculated in the weak coupling limit as \begin{equation} c=c_0\{1-\langle [\tau,\tau ]\rangle /2Mc_0^2\}, \end{equation} where $c_0$ is the sound velocity without electron-phonon coupling, $M$ is the ion mass, and for clarity we have suppressed the indeces both for the stress tensor component and for the sound velocity. For a highly anisotropic ($t_a\gg t_b\gg t_c$) tight-binding dispersion \begin{equation} \varepsilon({\bf p})=-2t_a\cos(ap_x)-2t_b\cos(bp_y) -2t_c\cos(cp_z)-\mu, \end{equation} widely used for CDW and SDW materials\cite{Yam}, the velocity and stress tensor components are easily obtained, and their values on the open Fermi surface can conveniently be expressed by the component of ${\bf p}$ perpendicular to the chains (${\bf x}$ direction). However, since the Green`s functions for CDW and SDW are usually written in the left-right spinor representation\cite{VM1}, involving measuring momenta from $\pm{\bf Q}/2$ with the density wave wavevector ${\bf Q}=(2p_F,\pi/b,\pi/c)$, we should express the stress tensor elements in a compatible manner. It turns out, that for each stress tensor component the term proportional to the unit matrix dominates, therefore we have \begin{eqnarray} \tau_{xx}=&mv_F^2\{1-[t_b/t_a\sin(ap_F)]^2 \sin^2(bp_y)\}\nonumber\\ \tau_{yy}=&mv_y^2[1+\cos(2bp_y)]\nonumber\\ \tau_{xy}=&mv_Fv_y\sqrt{2}\cos(bp_y). \end{eqnarray} Here $p_F$ is the Fermi momentum, $v_F=2at_a\sin(ap_F)$ is the Fermi velocity in the chain direction, while $v_y=\sqrt{2}bt_b\ll v_F$ is a typical velocity in the perpendicular direction. In the followings we restrict our study to the $x-y$ plane, since behavior involving the $z$ direction should be similar to that of the $y$ direction. We note, that all components of the stress tensor depend on momentum through the combination $\varphi=bp_y$ only. \section{Pinned density waves} In this section we consider sound propagation with no applied electric field. The density wave is pinned, therefore the condensate is unable to contribute to correlation functions, including that for the stress tensor. Mathematically this situation can be simulated by setting the coupling of the stress tensor (and of the density) to the phason to zero. Then the stress tensor couples only to the density fluctuations, resulting in the well known Coulomb screening\cite{KF}: \begin{equation} \langle [\tau,\tau]\rangle=\langle[\tau,\tau]\rangle_0- {\langle[\tau,n]\rangle_0\langle[n,\tau]\rangle_0\over q^2/4\pi e^2+\langle[n,n]\rangle_0}. \end{equation} Here $n$ stands for the electronic particle density, and $\langle[A,B]\rangle_0$ denotes a correlation function, in which only the effect of impurity scattering is taken into account. The density correlator $\langle[n,n]\rangle_0$ in the presence of impurity scattering was evaluated in \cite{VM1}. A straightforward extension of that calculation confirms that under the circumstances of the sound experiment ($lq\ll 1$, where $l$ is the mean free path) the stress tensor correlator has two distinct contributions: \begin{equation} \langle[\tau,\tau]\rangle_0=\langle\tau(\varphi)\rangle_{\varphi}^2 \langle[n,n]\rangle_0+\langle[\delta\tau(\varphi)]^2\rangle_{\varphi} \langle[n,n]\rangle_0^{no vertex}. \end{equation} The first contribution features only the average of the stress tensor $\langle\tau(\varphi)\rangle_{\varphi}= (2\pi)^{-1}\int_{-\pi}^\pi d\varphi\tau(\varphi)$, and is proportional to the diffusive (vertex corrected) density correlator. The second contribution containes only the fluctuating part of the stress tensor component $\delta\tau(\varphi)=\tau(\varphi)-\langle\tau(\varphi) \rangle_{\varphi}$, and therefore it is proportional to the density correlator $\langle[n,n]\rangle_0^{no vertex}$ calculated without vertex corrections. According to the same argument, only the average of the stress tensor couples to the density, therefore \begin{equation} \langle[\tau,n]\rangle_0= \langle[n,\tau]\rangle_0=\langle\tau(\varphi)\rangle_{\varphi} \langle[n,n]\rangle_0. \end{equation} Combining Eqs.(8)-(10) we see that in the long wavelength limit appropriate for the sound experiment ($q\approx 1/L$, where $L$ is the sample size), the average part of the stress tensor ($s$-wave component, proportional to density) is completely screened out by the Coulomb interaction, and only the fluctuating part contributes to the correlation function: \begin{equation} \langle[\tau,\tau]\rangle=\langle[\delta\tau(\varphi)]^2 \rangle_{\varphi}\langle[n,n]\rangle_0^{no vertex}. \end{equation} The situation here is the same as in the electronic Raman scattering, where the long-range Coulomb interaction suppresses the density (charge) fluctuations, and only non $s$-wave channels survive\cite{AG}. The evaluation of $\langle[n,n]\rangle_0^{no vertex}$ can be done starting with the results of \cite{VM1}. The calculation is rather technical, therefore we delegate it to the Appendix, and we give here the results only. Without vertex corrections there is no diffusion pole, and both the wavenumber ${\bf q}$ and the frequency $\omega$ could be set to zero. However we keep a finite (but small) frequency for finite imaginary part of the correlator: \begin{equation} \langle[n,n]\rangle_0^{no vertex}=N_F{i\tilde\Gamma_{qp}(1-\tilde f) \over \omega+i\tilde\Gamma_{qp}}, \end{equation} where $N_F$ is the density of states at the Fermi surface. The corresponding "unrenormalized" condensate density $\tilde f$ (for the general formula see Eq.(32) in the Appendix) and quasiparticle damping $\tilde\Gamma_{qp}$ are evaluated in two limiting cases, close to $T_c$ and close to zero temperature as \begin{eqnarray} \tilde f(T\rightarrow T_c)=&-2({\Delta\over 4\pi T})^2 \psi^{\prime\prime}({1\over 2}+{\Gamma\over 2\pi T})\approx 7\zeta(3)({\Delta\over 2\pi T})^2\nonumber\\ \tilde f(T\rightarrow 0)=&1-3\pi\alpha/16, \end{eqnarray} and \begin{eqnarray} \tilde\Gamma_{qp}(T\rightarrow T_c)=&2\Gamma\nonumber\\ \tilde\Gamma_{qp}(T\rightarrow 0)=&{9\pi\over 32}{\Delta \alpha^{8/3}\over u_0^2(5-4\alpha^{2/3})}{G\over T}e^{G/T}. \end{eqnarray} Here $\Delta$ is the density wave order parameter, $\Gamma= \Gamma_F+\Gamma_B/2$ is a combination of the impurity forward and backscattering rate, $\alpha=\Gamma/\Delta$, $u_0^2=1- \alpha^{2/3}$ and $G=\Delta u_0^3$ is the density wave gap. As it is seen from Eq.(13), $\tilde f$ increases linearly in $(T_c-T)$ below $T_c$, but it is slightly less than one at $T=0$ ($\Gamma$ is usually an order of magnitude smaller than $T_c$). Nevertheless, the temperature dependence of the sound velocity in the pinned case will still be qualitatively the same as in the collisionless limit\cite{MV1}. The relative change of the sound velocity compared to the normal state ($c_n$) is easily obtained from Eqs.(5), (11) and (12) as \begin{equation} (c-c_n)/c_0=\lambda\tilde f, \end{equation} where the effective coupling \begin{equation} \lambda={N_F\over 2Mc_0^2}\langle[\delta\tau(\varphi)]^2 \rangle_{\varphi}. \end{equation} Using Eq.(7), these effective couplings for the various sound waves are \begin{eqnarray} \lambda_{xx}=&{N_F(mv_F^2)^2\over 16Mc_0^2}[{t_b\over t_a\sin( ap_F)}]^4\nonumber\\ \lambda_{yy}=&{N_F(mv_y^2)^2\over 4Mc_0^2}\nonumber\\ \lambda_{xy}=&{N_F(mv_Fv_y)^2\over 2Mc_0^2}. \end{eqnarray} Since $v_y/v_F\approx t_b/t_a\approx 1/10$ in many quasi one dimensional materials, we expect that the relative increase of the sound velocity below $T_c$ will be a factor $10^2$ smaller for longitudinal sound than for transverse sound. \section{Electromechanical effect} If an external electric field in excess of the threshold field $E_T$ of the nonlinear conductivity is applied in the chain direction, then the condensate is depinned and is able to contribute to various correlation functions\cite{MV1}. The best known example is of course the conductivity itself, but the situation is the same for the stress tensor correlator as well. The collective contribution to $\langle[\tau,\tau]\rangle$ can be obtained if we allow both the stress tensor and the density to couple to the phason (in the previous section this coupling was blocked due to pinning).In this case the stress tensor correlator has another contribution $\langle[\tau,\tau]\rangle^{coll}$ in addition to the one calculated in the previous section, namely: \begin{equation} \langle[\tau,\tau]\rangle^{coll}={U\langle[\tau,\delta\Delta] \rangle^{Coul}\langle[\delta\Delta,\tau]\rangle^{Coul}\over 1-U\langle[\delta\Delta,\delta\Delta]\rangle^{Coul}}. \end{equation} Here $U$ is the on-site Coulomb repulsion responsible for the formation of the SDW state (for CDW it should be replaced by the phonon propagator, but that does not affect our conclusions), $\delta\Delta$ is the phase fluctuation of the order parameter, and $\langle[A,B]\rangle^{Coul}$ is the correlation function of quantities $A$ and $B$ including the effect of the long-range Coulomb interaction (like in Eq.(8)). As we have seen earlier, in the long wavelength limit this yields: \begin{equation} \langle[A,B]\rangle^{Coul}=\langle[A,B]\rangle_0-\langle[A,n] \rangle_0\langle[n,B]\rangle_0/\langle[n,n]\rangle_0. \end{equation} First we consider if the allowed coupling to the phason actually takes place for various sound waves. According to Eq.(18) we need to examine $\langle[\tau,\delta\Delta]\rangle^{Coul}$, which is given by Eqs.(19) and (10) as \begin{equation} \langle[\tau,\delta\Delta]\rangle^{Coul}=\langle[\tau,\delta \Delta]\rangle_0-\langle\tau(\varphi)\rangle_{\varphi}\langle [n,\delta\Delta]\rangle_0. \end{equation} The density-phason correlator $\langle[n,\delta\Delta]\rangle_0$ was evaluated in \cite{VM1}. Here we only reiterate that result in the limit of experimental interest $lq\ll c_0/v_F$ (dynamic limit): \begin{equation} \langle[n,\delta\Delta]\rangle_0=iN_F{\langle\zeta(\varphi) \rangle_{\varphi}\over 2\Delta}f_d, \end{equation} where in our two dimensional geometry the wavenumber ${\bf q}$ appears in $\zeta(\varphi)=v_Fq_x+\sqrt{2}v_yq_y\cos\varphi$, and the condensate density in the dynamic limit is given by: \begin{eqnarray} f_d(T\rightarrow T_c)=&{\Delta^2\over 2\pi T\Gamma_B} \psi^\prime({1\over 2}+{\Gamma\over 2\pi T})\approx {\pi\Delta^2\over 4T\Gamma_B}\nonumber\\ f_d(T\rightarrow 0)=&1. \end{eqnarray} We recall that the same $f_d$ appears in the current-phason correlator $\langle[j,\delta\Delta]\rangle_0$, as well as in the phason propagator\cite{MV2}. Note that $f_d$ increases from zero much faster below $T_c$ than $\tilde f$ does, and that at zero temperature it saturates exactly to $1$. The stress tensor-phason correlator can be calculated similarly: \begin{equation} \langle[\tau,\delta\Delta]\rangle_0=iN_F{\langle\tau(\varphi) \zeta(\varphi)\rangle_{\varphi}\over 2\Delta}f_d. \end{equation} Now we shall examine Eq.(20) for different sound waves in order to determine if there is a collective contribution to the corresponding stress tensor correlator. We consider longitudinal and transverse sound waves propagating in the ${\bf x}$ and ${\bf y}$ directions. Clearly, the second (screening) term in Eq.(20) is nonzero only for the longitudinal sound propagating in the chain direction (${\bf q}\parallel {\bf u} \parallel {\bf x}$), in which case it completely cancels the first term, leading to no collective contribution. The other longitudinal sound propagating perpendicular to the chains (${\bf q}\parallel {\bf u}\parallel {\bf y}$) does not couple to the phason either, because $\langle\tau_{yy}(\varphi)\cos\varphi\rangle_{\varphi} =0$ (see Eq.(7)). The coupling of the transverse wave propagating in the chain direction (${\bf q}\parallel {\bf x}$ and ${\bf u}\parallel {\bf y}$) is controlled by $\langle \tau_{xy}(\varphi)\rangle_{\varphi}=0$, yielding again no collective contribution. This means that in all of the above three cases there will be no electromechanical effect. For the rest of this section we will concentrate on the only interesting case, when the transverse sound propagates perpendicular to the chains (${\bf q}\parallel {\bf y}$ and ${\bf u}\parallel {\bf x}$). In this case there will be coupling to the phason, since \begin{equation} \langle[\tau_{xy},\delta\Delta]\rangle^{Coul}= \langle[\tau_{xy},\delta\Delta]\rangle_0= iN_F{f_d\over 2\Delta}mv_Fv_y^2q_y \end{equation} is nonzero. Now we have to consider the denominator in Eq.(18). Since $\langle[n,\delta\Delta]\rangle_0=0$ for $q_x=0$, therefore $\langle[\delta\Delta,\delta\Delta]\rangle^{Coul}= \langle[\delta\Delta,\delta\Delta]\rangle_0$, and we can use the result for the phason propagator calculated in \cite{VM1}, which in our case reduces to \begin{equation} 1-U\langle[\delta\Delta,\delta\Delta]\rangle_0= {UN_Ff_d\over (2\Delta)^2}[(v_yq_y)^2-i\omega\Gamma_{ph}]. \end{equation} Here $\Gamma_{ph}$ is the phason damping rate and is given by \begin{eqnarray} \Gamma_{ph}(T\rightarrow T_c)=&2\Gamma_B\nonumber\\ \Gamma_{ph}(T\rightarrow 0)=&{8u_0^2T\Gamma_B\over 3\alpha^{4/3}G}e^{-G/T}. \end{eqnarray} The phason damping freezes out for low temperature, and approaches $2\Gamma_B$ at $T_c$. Note the discontinuity in $\Gamma_{ph}$ at $T_c$ (approaching from above $\Gamma_{ph} \approx 2\pi^3T/7\zeta(3)$), which is the consequence of the finite order parameter $\Delta$ below $T_c$ exceeding almost immediately the energy scale set by $\omega$ and $vq$. We are now able to write down the total correlation function for this sound wave in the unpinned case. Using Eqs.(11),(18), (24) and (25) we obtain \begin{eqnarray} \langle[\tau_{xy},\tau_{xy}]\rangle=&N_F(mv_Fv_y)^2 \times\nonumber\\ \times&\left [ {i\tilde\Gamma_{qp}(1-\tilde f)\over \omega+i\tilde \Gamma_{qp}}+{(v_yq_y)^2f_d\over(v_yq_y)^2-i\omega\Gamma_{ph}} \right ]. \end{eqnarray} The above equation means that the collective contribution (the second term on the right hand side) due to the moving condensate recovers some of the screening of the ion motion lost because of the decrease in the number of quasi particles. In fact, at zero temperature it overcompensates somewhat, since $f_d>\tilde f$. Therefore we expect the electromechanical effect for the transverse sound polarized in the chain direction only. According to Eq.(27), close to $T_c$ the electromechanical effect on the sound velocity should be much smaller than the temperature effect, while at low temperatures the softening is somewhat bigger than the hardening was upon cooling. Although the collective contribution in Eq.(27) does have a frequency dependence, it does not appear to describe the suppresion of the electromechanical effect when the frequency is increased\cite{Xi3}. This is rather puzzling, although the above results for sound propagation may not translate literally to the flexural experiment. \section{Conclusions} We have derived for the first time the appropriate stress tensor for quasi-one dimensional electron systems. Under conditions of a sound velocity measurement ($lq\ll 1$) the long-range Coulomb interaction has a simple role to suppress the $s$-wave channel as in the theory of electronic Raman scattering. In highly anisotropic systems like Bechgaard salts the increase of the sound velocity in the density wave state is two orders of magnitude smaller for longitudinal sound waves than for transverse ones. We also find, that a sound wave with polarization perpendicular to the chain direction can not couple to the phason, because the density wave condensate can only move parallel to the chains. The coupling of the longitudinal sound propagating in the chain direction is screened away by the Coulomb interaction, which leaves us only the transverse sound wave propagating perpendicular to the chains as the one which does couple to the phason, and shows the electromechanical effect. Recently Britel {\it et.al.} measured the elastic constant $c_{44}$ of (Ta$_{1-x}$Nb$_x$Se$_4$)$_2$I at 15MHz in the geometry ${\bf u}\parallel {\bf x}$\cite{Mon}, and found a relative reduction of order $10^{-4}$ in the presence of an electric field approximately $10E_T$. This seems to be consistent with our analysis, although the observed effect appears to be a little too small. \section*{Acknowledgments} This publication is sponsored by the U.S.-Hungarian Science and Technology Joint Fund in cooperation with the National Science Foundation and the Hungarian Academy of Sciences under Project No. 264/92a, which enabled one of us (A.V.) to enjoy the hospitality of USC. This work is also supported in part by the National Science Foundation under Grant No. DMR92-18317 and by the Hungarian National Research Fund under Grants No. OTKA15552 and T4473. \section*{Appendix} We evaluate here the density correlation function $\langle[n,n] \rangle_0^{no vertex}$ (Eqs.(12)-(14) in the text). We start with the corresponding thermal product (See \cite{VM1}): \begin{equation} \langle[n,n]\rangle_0^{no vertex}=N_F[1-\pi T\sum_n F(iu_n,iu_n^\prime)], \end{equation} where $N_F$ is the density of states, and $u_n$ and $u_n^\prime$ are related to the Matsubara frequencies $\omega_n$ and $\omega_n^\prime= \omega_{n-\nu}$ by \begin{equation} \omega_n/\Delta=u_n[1-\alpha(u_n^2+1)^{-1/2}], \end{equation} with $\Delta$ the order parameter, $\alpha=\Gamma/\Delta$ and $\Gamma=\Gamma_F+\Gamma_B/2$ a combination of the impurity forward and backscattering rates. Neglecting vertex corrections in the relevant formulas in \cite{VM1} leads to: \begin{equation} F(u,u^\prime)={1+{1+uu^\prime\over (1-u^2)^{1/2}(1-u^{\prime 2}) ^{1/2}}\over \Delta[(1-u^2)^{1/2}+(1-u^{\prime 2})^{1/2}]}, \end{equation} where $u$ and $u^\prime$ are analytic continuations of $iu_n$ and $iu_n^\prime$, and in the absence of the diffusion pole the wavenumber ${\bf q}$ was already set to zero. While evaluating the correlation function we follow the standard method\cite{Zit}. Expanding up to linear order in $\omega$ we obtain \begin{equation} \langle[n,n]\rangle_0^{no vertex}=N_F(1-\tilde f+i\omega I), \end{equation} where \begin{equation} \tilde f={\pi T\over\Delta}\sum_n(u_n^2+1)^{-3/2}, \end{equation} and \begin{eqnarray} I=&{1\over 2\Delta}\int_G^\infty{dE\over 2T}\cosh^{-2}\left ({E\over 2T}\right )\times\nonumber\\ \times&\left [{h^\prime\over {\rm Re}(1-u^2)^{1/2}}-{\rm Re} (1-u^2)^{-3/2}\right ]. \end{eqnarray} Here $h^\prime=(1/2)[1+(|u|^2+1)/|u^2-1|]$, and $G=\Delta u_0^3$ is the gap with $u_0^2=1-\alpha^{2/3}$. The above equations can be evaluated in two limiting cases, with temperature close to $T_c$ and close to zero, and can be brought to the form of Eqs.(12)-(14) in the text.
\section{Introduction} For an agent\footnote{Unless otherwise indicated, we will use the term ``agent'' to refer to both individual agents and sets of agents.} to be able to perform an action, it must satisfy both the physical and knowledge preconditions of that action \cite{Moore:85,Morgenstern:87}. For example, for an agent to pick up a particular tower of blocks, it must (1)~know how to pick up towers in general, (2)~be able to identify the tower in question, and (3)~ have satisfied the (physical) preconditions or constraints associated with picking up towers (e.g., it must have a free hand). These conditions must hold whether the agent is planning an action on its own or is involved in a collaborative planning effort with other agents. In this paper, we provide an axiomatization of knowledge preconditions for the SharedPlan model of collaborative activity \cite{Grosz-Sidner:90,LGS:90,Grosz-Kraus:93}. This model draws upon past work \cite{Moore:85,Morgenstern:87}, but adapts it to the collaborative situation. We briefly describe the SharedPlan framework in Section~\ref{sp}, and then, in Section~\ref{know-prec}, present our axiomatization of knowledge preconditions. In Section~\ref{role}, we demonstrate the use of knowledge preconditions in accounting for information-seeking subdialogues, such as those in Figure~\ref{ac}. We then compare our approach to the alternative accounts \cite{Litman-Allen:87,Lambert-Carberry:91,Ramshaw:91}. \begin{small} \begin{figure}[hbtp] \hspace*{0.2in} \psfig{figure=knowp-d.eps,width=3.1in} \caption{\small Information-Seeking Subdialogues [Grosz, 1974]\label{ac}} \end{figure} \end{small} \vspace*{-3ex} \section{SharedPlans} \label{sp} The SharedPlan formalism is a mental-state model of collaborative plans with roots in Pollack's~\shortcite{Pollack:90} work on single-agent plans. For a group of agents $GR$ to have a {\em full SharedPlan} (FSP) for an act $\alpha$, they must satisfy the requirements given in Figure~\ref{sp-reqts}. When the agents have satisfied only a subset of these requirements, they are said to have a {\it partial SharedPlan} (PSP).\footnote{This description of a PSP is only a rough, though useful, approximation to the formal definition given by Grosz and Kraus~\shortcite{Grosz-Kraus:93}.} The bracketed terms in Figure~\ref{sp-reqts} indicate the operators used by Grosz and Kraus~\shortcite{Grosz-Kraus:93} to formalize each requirement. Requirement~(1) in Figure~\ref{sp-reqts} refers to the agents' {\em recipe} \cite{Pollack:90} for $\alpha$. Recipes are modeled in Grosz and Kraus's definitions as sets of constituent acts and constraints. To perform an act $\alpha$, an agent must perform each constituent act in $\alpha$'s recipe according to the constraints of that recipe. Actions themselves may be further decomposed into act-types and parameters. We will represent an action $\alpha$ as a term of the form $\bar{\alpha}(p_1,\ldots,p_n)$ where $\bar{\alpha}$ represents the act-type of the action and the $p_i$ its parameters. \begin{figure}[t] \begin{small} For a group of agents $GR$ to have an FSP for $\alpha$ \begin{enumerate} \item $GR$ must have mutual belief of a {\em recipe} for $\alpha$ \item For each single-agent constituent act of the recipe, there must be an agent $G_{\beta_i} \in GR$, such that \begin{enumerate} \vspace*{-1ex} \item $G_{\beta_i}$ intends to perform $\beta_i$ {\bf [Int.To]} \item $G_{\beta_i}$ believes that it can perform $\beta_i$ {\bf [BCBA]} \item $G_{\beta_i}$ has a full individual plan for $\beta_i$ {\bf [FIP]} \item The group $GR$ has mutual belief of (2a)-(2c) \item Each member of $GR$ intends that $G_{\beta_i}$ succeed {\bf [Int.Th]} \end{enumerate} \item For each multi-agent constituent act of the recipe, there must be a subgroup of agents $GR_{\beta_i} \subseteq GR$ such that \begin{enumerate} \item $GR_{\beta_i}$ mutually believe that they can perform $\beta_i$ {\bf [MBCBAG]} \item $GR_{\beta_i}$ has a full SharedPlan for $\beta_i$ {\bf [FSP]} \item The group $GR$ has mutual belief of (3a)-(3b) \item Each member of $GR$ intends that $GR_{\beta_i}$ succeed {\bf [Int.Th]} \end{enumerate} \end{enumerate} \vspace*{-2ex} \caption{\small FSP Requirements \label{sp-reqts}} \end{small} \end{figure} \section{Knowledge Preconditions} \label{know-prec} Grosz and Kraus~\shortcite{Grosz-Kraus:93} use the operators BCBA (read ``believes can bring about'') and MBCBAG (read ``mutually believe can bring about group'') to formalize respectively requirements~(2b) and~(3a) in Figure~\ref{sp-reqts}. Although these operators are intended to specify the conditions under which an agent is able to perform an action, their definitions explicitly require only that an agent satisfy the physical preconditions or constraints associated with an action to be able to perform it. Because an agent is not truly capable of performing an act unless it possesses the appropriate knowledge, the definitions of BCBA and MBCBAG must be augmented with an axiomatization of knowledge preconditions. The following observations made by Morgenstern~\shortcite{Morgenstern:87}, but recast in our terminology, must be represented in such an axiomatization: \begin{enumerate} \vspace*{-1ex} \item Agents need to know recipes for the acts they perform. \vspace*{-1.5ex} \item All agents have some primitive acts in their repertoire. \vspace*{-1.5ex} \item Agents must be able to identify the parameters of the acts they perform. \vspace*{-1.5ex} \item Agents may know only some descriptions of an act. \vspace*{-3.5ex} \item Agents know that the knowledge necessary for complex acts derives from that necessary for their component acts. \vspace*{-1.5ex} \end{enumerate} Our axiomatization of knowledge preconditions is based on Morgenstern's observations, but adapted to the requirements of individual and shared mental-state plans.\footnote{A comparison of our formalization with those of Morgenstern~\shortcite{Morgenstern:87} and Moore~\shortcite{Moore:85} can be found elsewhere \cite{KEL:94}.} We use the predicates {\em has.recipe\/} and {\em id.params\/} to represent explicitly observations~(1) and (3) above. The remaining observations are implicitly represented by the way in which these two knowledge precondition relations are defined. Observation~(2) is modeled as the base case of {\em has.recipe}, and observation~(5) is modeled by the use of {\em has.recipe\/} within the recursive plan definitions. Observation~(4) requires that the knowledge precondition relations be intensional, rather than extensional; within their scope it should not be possible to freely substitute one representation of an action for another. We thus define {\em has.recipe\/} and {\em id.params\/} to hold of action {\em descriptions}, rather than actions. Action descriptions are intensional objects; one action description can be substituted for another only if the descriptions are the same. For example, although $555$-$1234$ and $phone$-$number(speech$-$lab)$ may be extensionally equivalent, the descriptions $\rule[1.2ex]{.05em}{.24em}^{\hspace*{-0.05em}\rule[.6ex]{.25em}{.05em}}\!\,{555}$-${1234}\:\rule[1.2ex]{.05em}{.24em}^{\hspace*{-.26em}\rule[.6ex]{.25em}{.05em}}$ and $\rule[1.2ex]{.05em}{.24em}^{\hspace*{-0.05em}\rule[.6ex]{.25em}{.05em}}\!{phone}$-${number(speech}$-${lab)}\:\rule[1.2ex]{.05em}{.24em}^{\hspace*{-.26em}\rule[.6ex]{.25em}{.05em}}$ are not. By convention, we will omit the corner quote notation in what follows and simply take the appropriate arguments of the predicates to represent action descriptions rather than actions. Although Morgenstern's observations are most naturally expressed informally in terms of knowledge, we formalize them using belief to allow for the possibility of an agent's being incorrect. Although it is true that an agent cannot successfully act unless its beliefs about recipes and parameters are correct, having to {\it know} the recipes and parameters is too strong a requirement for collaborating agents \cite{KEL:94}. \subsection{Determining Recipes: {\em has.recipe}} \label{recipe} For an agent to be able to perform an act $\alpha$, it must know {\em how\/} to perform $\alpha$; i.e., it must have a recipe for the act. The relation $has.recipe(G,\alpha,R,T)$ is used to represent that agent $G$ has a recipe $R\,$ for an act $\alpha$ at time $T$. It is formalized as follows: \begin{footnotesize} \begin{tabbing} {}~~~~~~~~$has.recipe(G,\alpha,R,T) \Leftrightarrow$\\ {}~~~~~~~~(1)~~~~$[$\=$basic.level(\alpha)~~\wedge$\\ \>$BEL(G,basic.level(\alpha),T) \wedge R=R_{Empty}]~~\vee$ \\ {}~~~~~~~~(2)~~~~$[\neg basic.level(\alpha)~~\wedge$\\ {}~~~~~~~~(2a) \>$R=\{\beta_i,\rho_j\}~~\wedge$\\ {}~~~~~~~~(2a1) \>~~~~$\{$\=$[|G| = 1 \wedge BEL(G,R \in Recipes(\alpha),T)]~~\vee$\\ {}~~~~~~~~(2a2) \>\>$[|G| > 1 \wedge MB(G,R \in Recipes(\alpha),T)]\}]$ \end{tabbing} \end{footnotesize} Clause~(1) of the definition models Morgenstern's second observation, namely that agents do not need a recipe to perform a {\em basic-level\/} action, i.e., one executable at will \cite{Pollack:90}.\footnote{Basic-level actions are by their nature single-agent actions.} For non-basic-level actions (Clause~(2)), the agent of $\alpha$ (either a single agent (2a1) or a group of agents (2a2)) must believe that some set of acts, $\beta_i$, and constraints, $\rho_j$, constitute a recipe for $\alpha$. \subsection{Identifying Parameters: {\em id.params}} \label{params} An agent must also be able to identify the parameters of an act $\alpha$ to be able to perform it. For example, if an agent is told {\it ``remove the flywheel,''} as in the dialogue of Figure~\ref{ac}, the agent must be able to identify the flywheel in question. The relation $id.params(G,\alpha,T)$ is used to represent that agent $G$ can identify the parameters of act $\alpha$ at time $T$. If $\alpha$ is of the form $\bar{\alpha}(p_1,...,p_n)$, then $id.params(G,\alpha,T)$ is true if $G$ can identify each of the $p_i$. To do so, $G$ must have a description of each $p_i$ that is suitable for $\bar{\alpha}$. The relation {\em id.params\/} is defined as follows: \begin{footnotesize} \begin{tabbing} {}~~~~$id.params(G,\bar{\alpha}(p_1,\ldots,p_n),T) \Leftrightarrow$\\ {}~~~~~~~$(\forall i, 1 \leq i \leq n)~has.sat.descr(G,p_i,{\cal F}(\bar{\alpha},p_i),T)$ \end{tabbing} \end{footnotesize} The ability to identify an object is highly context dependent. For example, as Appelt points out~\shortcite[pg. 200]{Appelt:85}, ``the description that one must know to carry out a plan requiring the identification of `John's residence' may be quite different depending on whether one is going to visit him, or mail him a letter.'' The function ${\cal F}$ in the above definition is an oracle function intended to model the context-dependent nature of parameter identification. This function returns a suitable {\em identification constraint\/} \cite{Appelt-Kronfeld:87} for a parameter $p_i$ in the context of an act-type $\bar{\alpha}$. For example, in the case of sending a letter to John's residence, the constraint produced by the oracle function would be that John's residence be described by a postal address. The relation $has.sat.descr(G,P,C,T)$ holds of an agent $G$, a parameter description $P$, an identification constraint $C$, and a time $T$, if $G$ has a suitable description, as determined by $C$, of the object described as $P$ at time $T$. To formalize this relation, we utilize Kronfeld's~\shortcite{Kronfeld:86} notion of an individuating set. An agent's individuating set for an object is a maximal set of terms such that each term is believed by the agent to denote that object. For example, an agent's individuating set for John's residence might include its postal address as well as an identifying physical description such as ``the only yellow house on Cherry Street.'' To model individuating sets, we introduce a function $IS(G,P,T)$; the function returns an agent $G$'s individuating set at time $T$ for the object that $G$ believes can be described as $P$. This function is based on similar elements of the formal language that Appelt and Kronfeld~\shortcite{Appelt-Kronfeld:87} introduce as part of their theory of referring. The function returns a set that contains $P$ as well as the other descriptions that $G$ has for the object that it believes $P$ denotes. For an agent to suitably identify a parameter described as $P$, the agent must have a description, $P^\prime$, of the parameter such that $P^\prime$ is of the appropriate sort. For example, for an agent to visit John's residence, it is not sufficient for the agent to believe that the description ``John's residence'' refers to the place where John lives. Rather, the agent needs another description of John's residence, one such as ``the only yellow house on Cherry Street,'' that is appropriate for the purpose of visiting him. To model an agent's ability to identify a parameter (described as $P$) for some purpose, we thus require that the agent have an individuating set for the parameter that contains a description $P^\prime$ such that $P^\prime$ satisfies the identification constraint that derives from the purpose. The definition of {\em has.sat.descr\/} is thus as follows:\footnote{A more precise account of what it means to be able to identify an object is beyond the scope of this paper; for further details, see the discussions by Hobbs~\shortcite{Hobbs:85}, Appelt and Kronfeld~\shortcite{Appelt:85,Kronfeld:86,Appelt-Kronfeld:87}, and Morgenstern~\shortcite{Morgenstern-thesis}.} \begin{footnotesize} \begin{tabbing} {}~~~~$has.sat.descr(G,P,C,T) \Leftrightarrow$\\ {}~~~~~~$\{$\=$[$\=$|G| = 1~\wedge$\\ \>\>~~$(\exists P^\prime) BEL(G,[$\=$P^\prime \in IS(G,P,T)~~\wedge$\\ \>\>\>${\it suff\!.for.id}(C,P^\prime)],T)]~~\vee$\\[0.5ex] \>$[|G| > 1~\wedge$\\ \>\>~~$(\exists P^\prime) MB(G,(\forall G_j \in G) [$\=$P^\prime \in IS(G_j,P,T)~~\wedge$\\ \>\>\>${\it suff\!.for.id}(C,P^\prime)],T)]\}$ \end{tabbing} \end{footnotesize} The predicate ${\it suff\!.for.id}(C,P^\prime)$ is true if the constraint $C$ applies to the parameter description $P^\prime$. The oracle function ${\cal F}(\bar{\alpha},p_i)$ in {\em id.params\/} produces the appropriate identification constraint on $p_i$ given $\bar{\alpha}$. \section{The Role of Knowledge Preconditions in Language Processing} \label{role} We now show how the requirements of knowledge preconditions can be used in discourse processing. Our model of discourse processing is based on the theory of discourse structure proposed by Grosz and Sidner~\shortcite{Grosz-Sidner:86}. According to their theory, discourse structure consists of three interrelated components: a linguistic structure, an attentional state, and an intentional structure. The linguistic structure consists of discourse segments\footnote{The term discourse segment is a generalization of the term subdialogue. Whereas the term discourse segment applies to all types of discourse, the term subdialogue is reserved for segments that occur within dialogues.} and an embedding relationship among them; the bold rule in Figure~\ref{ac} indicates the linguistic structure of that discourse. Attentional state is an abstraction of the discourse participants' focus of attention; it serves as a record of those entities that are salient at any point in the discourse. Intentional structure is comprised of discourse segment purposes and their interrelationships, particularly that of {\em dominance}. A discourse segment purpose, or DSP, is a Gricean-like~\shortcite{Grice:69} intention that leads to the initiation of a discourse segment. One DSP is dominated by another if the satisfaction of the first provides part of the satisfaction of the second. Intentional structure plays a central role in discourse processing; an agent's comprehension of the utterances in a discourse relies on the recognition of this structure \cite{Grosz-Sidner:86}. Grosz and Sidner~\shortcite{Grosz-Sidner:90} proposed SharedPlans to provide a basis for recognizing intentional structure. They argued that discourses are fundamentally collaborative, and hence that a model of {\it shared\/} plans provides a more appropriate basis for discourse processing than a model of single-agent plans. However, the connection between SharedPlans and intentional structure was never specified. \subsection{SharedPlans as Intentional Structure} We have developed a model of discourse processing that provides that connection \cite{KEL:94}. Figure~3 illustrates the role of SharedPlans in modeling intentional structure. Each segment of a discourse has an associated SharedPlan. The purpose of the segment is taken to be intention that (Int.Th \cite{Grosz-Kraus:93}) the discourse participants form that plan. This intention is held by the agent who initiates the segment. In what follows, we will refer to that participant as the initiating conversational participant or ICP; the other participant is the OCP \cite{Grosz-Sidner:86}. Dominance relationships between DSPs are modeled using {\it subsidiary} relationships between SharedPlans. One plan is subsidiary to another if the completion of the first plan contributes to the completion of the second. Subsidiary relationships are discussed in more detail in Section~\ref{recog} The utterances of a discourse are understood in terms of their contribution to the SharedPlans associated with the segments of the discourse. Those segments that have been completed at the time of processing an utterance have a full SharedPlan (FSP) associated with them (e.g., segment~(2) in Figure~3), while those that have not have a partial SharedPlan (PSP) (e.g., segments~(1) and~(3) in Figure~3). \psfig{figure=struct-ijcai.eps,width=3.37in} \begin{center} Figure 3: Modeling Intentional Structure \end{center} For each utterance of a discourse, an agent must determine whether the utterance begins a new segment of the discourse, contributes to the current segment, or completes it \cite{Grosz-Sidner:86}. For an utterance to begin a new segment, it must indicate the initiation of a subsidiary plan. This case is described in further detail below. For an utterance to contribute to the current segment, it must advance the partial SharedPlan associated with the segment towards completion. That is, it must establish one of the beliefs or intentions required for the discourse participants to have a full SharedPlan, but missing from their current partial SharedPlan. For an utterance to complete the current segment, it must indicate that the purpose of that segment has been satisfied. For that to be the case, the SharedPlan associated with the segment must be an FSP rather than a PSP. That is, all of the beliefs and intentions required of an FSP, as indicated in Figure~\ref{sp-reqts}, must have been established over the course of the segment. A detailed description of the implemented algorithms used in modeling each of these cases can be found elsewhere \cite{KEL:94}. Here, we focus on the use of knowledge preconditions in accounting for the initiation of information-seeking subdialogues. We use the dialogue in Figure~\ref{ac} as an example and assume the role of the Expert (participant ``E'') in analyzing the discourse.\footnote{For simplicity of exposition, we will take participant ``E'' to be female and participant ``A'' to be male.} The dialogue in Figure~\ref{ac} was extracted from a larger discourse in which the Expert and Apprentice (participant ``A'') are collaborating on removing the pump of an air compressor. We thus take the purpose of the larger discourse to be \begin{footnotesize} \begin{tabbing} {}~~~DSP$_1$=$Int.Th(e,FSP(\{a,e\},remove(pump(ac1),\{a\})))$\footnotemark\\ {}~~~~~~where $ac1$ represents the air compressor the agents\\ {}~~~~~~~~are working on. \end{tabbing} \end{footnotesize} \footnotetext{We have omitted the time parameters for simplicity of exposition.} \subsection{Accounting for the Initiation of New Discourse Segments} \label{recog} To make sense of an {\em utterance}, an agent must provide an explanation for it in the form of an answer to the question, ``Why did the speaker say that to me?'' \cite{Sidner-Israel:81}. An OCP must provide a similar explanation for an ICP's initiation of a new discourse segment. This explanation takes the form of an answer to the question ``Why does the ICP want to engage in a segment with purpose DSP$_j$ at this point in our discourse?''; i.e., ``How is DSP$_j$ related to what we were talking about before?'' Subsidiary relationships between SharedPlans provide the basis for modeling the OCP's reasoning. One plan is subsidiary to another if the completion of the first plan contributes to the completion of the second. The most basic example of this relationship occurs within the FSP definition itself. As indicated in Figure~\ref{sp-reqts}, a full plan for an act $\alpha$ includes full plans for each subact in $\alpha$'s recipe as components (requirements~(2c) and (3b)). The plans for the subacts thus contribute to the plan for $\alpha$ and are therefore subsidiary to it. Subsidiary relationships may also arise in response to the other requirements of the FSP definition. For example, as discussed in Section~\ref{know-prec}, the BCBA operator used to model requirement~(2b) specifies that to be able to perform an act $\alpha$, an agent must (1)~have a recipe for $\alpha$ ($has.recipe$), (2)~be able to identify the parameters of $\alpha$ ($has.sat.descr$), and (3)~have satisfied the constraints associated with performing $\alpha$. The first of these requirements provides an explanation for the first subdialogue in Figure~\ref{ac}. The purpose of this subdialogue is represented as\footnote{We describe a method for recognizing DSPs of this form elsewhere \cite{KEL:94}.} \begin{footnotesize} \begin{tabbing} {}~~~~~DSP$_2$=\=$Int.Th(a,FSP(\{a,e\},$\\ \>~~$Achieve(has.$\=$recipe(a,$\\[-0.5ex] \>\>$remove(flywheel(ac1),\{a\}),R))))$ \end{tabbing} \end{footnotesize} and can be glossed as ``the Apprentice intends that the agents collaborate on his obtaining a recipe for the act of removing the flywheel of the air compressor.'' To account for the Apprentice's initiation of this subdialogue, the Expert must determine the relationship of DSP$_2$ to the purpose of the agents' preceding discourse, namely DSP$_1$. In this case, the Expert can reason that the Apprentice wants to engage in the subdialogue to obtain a recipe for the act of removing the flywheel so that he will be able to perform that act as part of the agents' SharedPlan to remove the pump. The plan in DSP$_2$ is thus subsidiary to that in DSP$_1$ by virtue of a knowledge precondition requirement of the latter plan. Figure~\ref{kp1-graphic} illustrates this analysis. Each box in the figure corresponds to a discourse segment and contains the SharedPlan used to model the segment's purpose. The SharedPlans are labeled so as to be co-indexed with the DSPs discussed above. The arrows indicate subsidiary relationships between SharedPlans, as explained by the text that adjoins them. When plan Pj is subsidiary to plan Pi, DSP$_j$ is dominated by DSP$_i$. \begin{small} \begin{figure*}[bt] \hspace*{1in} \psfig{figure=knowp-ijcai2.eps,width=5.7in} \caption{Analysis of the First Subdialogue in Figure~\protect\ref{ac}\label{kp1-graphic}} \end{figure*} \end{small} The information represented within each SharedPlan in Figure~\ref{kp1-graphic} is separated into two parts. Those beliefs and intentions that have been established at the time of the analysis are shown above the dotted line, while those that remain to be established, but that are used in determining subsidiary relationships, are shown below the line. The index in square brackets to the right of each constituent indicates the FSP requirement from which the constituent arose. As indicated in Figure~\ref{kp2-graphic}, the initiation of the second subdialogue in Figure~\ref{ac} is explained similarly. This time, however, it is the need to identify parameters of acts (requirement~(2) above) that leads to the initiation of the subdialogue. In addition, the parameter in question is a parameter of an act in a subsidiary individual plan of the Apprentice's. \begin{small} \begin{figure*}[tb] \hspace*{1in} \psfig{figure=knowp-2-ijcai2.eps,width=4.9in} \caption{Analysis of the Second Subdialogue in Figure~\protect\ref{ac}\label{kp2-graphic}} \end{figure*} \end{small} \subsection{Discussion} In this paper, we have shown that information-seeking subdialogues may be explained on the basis of knowledge precondition requirements. Our account of such subdialogues fits within a general framework for discourse processing in which the purpose of a subdialogue is modeled using a SharedPlan and is related to the purposes of other subdialogues based on the requirements of the FSP definition. Elsewhere~\cite{KEL:94}, we show that correction and subtask subdialogues, among others, may also be accounted for in this manner. In contrast, alternative plan-based accounts of dialogue understanding introduce multiple types of plans to account for the utterances in a discourse. For example, Litman and Allen~\shortcite{Litman-Allen:87}, propose the use of two types of plans to model clarification and correction subdialogues: {\em discourse plans\/} and {\em domain plans}. Domain plans represent knowledge about a task, while discourse plans represent conversational relationships between utterances and plans. Litman and Allen provide operators for the following discourse plans: \begin{itemize} \vspace*{-1ex} \item INTRODUCE-PLAN: introduce a new plan for discussion \vspace*{-1ex} \item CONTINUE-PLAN: execute the next action in a plan \vspace*{-1ex} \item TRACK-PLAN: talk about the execution of an action \vspace*{-1ex} \item MODIFY-PLAN: introduce a new plan by modifying a previous one \vspace*{-1ex} \item CORRECT-PLAN: correct a plan \vspace*{-1ex} \item IDENTIFY-PARAMETER: identify a parameter of an action in a plan \end{itemize} Under our approach, the recognition of discourse plans is unnecessary. The {\em fact} that a speaker is using an utterance to, for example, introduce a plan, or track a plan, or identify a parameter, need not be explicitly recognized for the purposes of utterance interpretation. Furthermore, we would argue that such facts are not intended to be recognized (cf. Grice~\shortcite{Grice:69}). Rather, they simply fall out of recognizing the relationship of an utterance to the current discourse structure, i.e., the currently active SharedPlans. For example, INTRODUCE-PLAN corresponds to initiating a new discourse segment, CONTINUE- or TRACK-PLAN to contributing to the current segment, and IDENTIFY-PARAMETER to initiating a new segment to satisfy a $has.sat.descr$ knowledge precondition requirement. Although the initiation of a new SharedPlan corresponds to the initiation of a new discourse segment under our approach, it is the SharedPlan that must be recognized and not a discourse plan that refers to that SharedPlan. Lambert and Carberry~\shortcite{Lambert-Carberry:91} have extended Litman and Allen's approach by introducing a third type of plan. {\em Problem-solving plans}, such as BUILD-PLAN and INSTANTIATE-VARS, are used to model the process by which agents construct domain plans. Under our approach, the need to explicitly recognize problem-solving plans is also avoided. The fact that an agent is building a plan or instantiating a variable is a byproduct of understanding an utterance by relating it to the current discourse structure. BUILD-PLAN corresponds to initiating a new discourse segment to satisfy a $has.recipe$ knowledge precondition requirement, while INSTANTIATE-VARS corresponds to initiating one to satisfy a $has.sat.descr$ requirement. Unlike Lambert and Carberry's approach, however, and Litman and Allen's as well, our approach actually recognizes this structure. The other approaches are essentially utterance-to-utterance based and thus do not recognize discourse segments as separate units. Ramshaw~\shortcite{Ramshaw:91} has added a different third type of plan, {\em exploration plans}, to Litman and Allen's two types. Exploration plans are intended to model the process by which agents explore courses of actions. Although we have not yet incorporated such reasoning into our model, we hypothesize that the exploration of plans can be modeled, without the introduction of a new plan type, by reasoning about an agent's {\it potential intentions} and the process by which they become full-fledged intentions \cite{Grosz-Kraus:93,BIP:88}. These alternative approaches share an important property that distinguishes them from our approach; they take a {\em data-structure} view of plans, rather than a {\em mental phenomenon} view \cite{Pollack:90}. Whereas data-structure plans are essentially ``recipes-for-action,'' mental phenomenon plans are a ``structured collection of beliefs and intentions'' \cite[pg. 77]{Pollack:90}.\footnote{Although Lambert and Carberry~\shortcite{Lambert-Carberry:91} adopt Pollack's~\shortcite{Pollack:90} terminology in presenting their theory, their ``plans'' are not mental state plans in Pollack's sense.} Data-structure plans thus describe {\em what} an agent is doing with an utterance, but not {\em why} the agent is doing it. For example, although the constraints of Litman and Allen's IDENTIFY-PARAMETER discourse plan force the plan to be related to another plan that involves the parameter to be identified, IDENTIFY-PARAMETER does not explain why this information is desired; it does not capture that agents need to know parameters {\em to be able to} perform acts involving them.\footnote{Although Litman and Allen {\em describe} IDENTIFY-PARAMETER as providing ``a suitable description for a term instantiating a parameter of an action such that the hearer is then able to execute the action'' \cite[pg. 173]{Litman-Allen:87}, the IDENTIFY-PARAMETER operator itself does not include a {\em formalization} of the last condition, i.e., that the parameter description should enable the execution of the action.} It thus fails to model the essential knowledge precondition nature of identifying a parameter. Although it is possible to impose a mental phenomenon interpretation on top of a data-structure plan, doing so does not result in a mental phenomenon plan \cite{Pollack:90}. Saying that G1 {\em intends} to IDENTIFY a PARAMETER fails to address why G1 intends to do so. The need to explain an utterance is not unique to interpretation. Moore and Paris~\shortcite{Moore-Paris:93} have shown that a similar need exists in generation. In particular, they have argued that RST-based text plans must be augmented with intentional structure. Otherwise, a system has no record of why it said what it did and is thus unable to respond effectively if a hearer does not understand or accept its utterances. \section{Conclusion} In this paper, we have presented an axiomatization of knowledge preconditions for the SharedPlan model of collaborative activity \cite{Grosz-Kraus:93}. We have also shown how the requirements of knowledge preconditions can be used to account for information-seeking subdialogues in discourse. Our account of this phenomenon fits within a general framework for discourse processing in which SharedPlans and relationships among them are used to model the intentional component of Grosz and Sidner's~\shortcite{Grosz-Sidner:86} theory of discourse structure. Unlike the alternative approaches, our approach recognizes and makes use of discourse structure. In addition, it does not require the introduction of new plan types. \begin{small}
\section{Introduction} Since the wormhole-based time machine was proposed \cite{Tho} much efforts have been directed towards finding a mechanism that could "protect causality" and destroy such a time machine. One of the most popular ideas is that the creation of the time machine might be prevented by quantum effects since as it is claimed in \cite{KimT} "at any event in spacetime, which can be joined to itself by a closed null geodesic, the vacuum fluctuations of a massless scalar field should produce a divergent renormalized stress-energy tensor". The considerations leading to such a claim I shall call hereafter "FKT approach". \par In essence, the FKT approach amounts to the following \cite{Fro,KimT} (see also \cite{Kli}). The vacuum expectation value of the stress-energy tensor $\langle T_{\mu\nu}\rangle$ of the field $\phi$ in the (multiply connected) spacetime $M$ containing a time machine is found by applying some differential operator $D_{\mu\nu}$ to the Hadamard function \begin{equation} \label{DefG} G^{(1)}(X,X')\equiv\bigl\langle\{\phi(X),\phi(X')\}\bigr\rangle. \end{equation} To find $G^{(1)}$ it is proposed to use the formula \begin{equation} \label{Cov} G^{(1)}(X,X^{\prime})=G^\Sigma\equiv \sum_n{\widetilde G}^{(1)}(X, \gamma^n X^{\prime}) . \end{equation} Here $\widetilde G^{(1)}$ is the Hadamard function of $\phi$ in the spacetime $\widetilde M$, which is the universal covering space for $M$, and $\gamma^n X\in \widetilde M$ is the $n$-th inverse image of $X\in M$ ($\gamma^0X$ is identified in (\ref{Cov}) with $X$). The advantage of the use of (\ref{Cov}) is that $\widetilde G^{(1)}$ is supposed to have the Hadamard form: \begin{equation} \label{Had} {\widetilde G}^{(1)}(X,X')=\tilde u\sigma^{-1}+\tilde v\ln{|\sigma|} +\mbox{nonsingular terms} \end{equation} where $\sigma$ is half the square of the geodesic distance between $X$ and $ X'$, and $\tilde u,\ \tilde v$ are some smooth functions. We might think thus that \begin{equation} \begin{array}{lcl}\displaystyle{ \langle T_{\mu\nu}\rangle^{{\rm ren}}_{M\vphantom{\widetilde M}} }&= \displaystyle{\langle T_{\mu\nu}\rangle^{{\rm ren}}_{\widetilde M} + \sum _{n\not=0}\lim_ { X'\to X\atop{}} D_{\mu\nu}{\widetilde G}^{(1)} (X, \gamma^n X^{\prime})}\\ \strut&&\\&\to& \displaystyle{ \langle T_{\mu\nu}\rangle^{{\rm ren}}_{\widetilde M} + \sum _{n\not=0}\;\lim_ {X'\to X\atop X\to\, \mbox{\scriptsize horizon}} D_{\mu\nu} (\tilde u\sigma_n^{-1}+\tilde v\ln{|\sigma_n|}). }\end{array} \label{Tren} \end{equation} Here $\sigma_n\equiv \sigma(X,\gamma^nX')$ and the subscript "ren" (renormalized) has appeared because renormalization of $\langle {\bf T}\rangle_{\widetilde M}$ and of $\langle {\bf T}\rangle_{M\vphantom{\widetilde M}}$ requires substraction of the same terms. The last series in (\ref{Tren}) diverges (since $\sigma_n\to0$, when $X$ approaches the horizon), so the conclusion is made that the appearance of a closed timelike curve must be prevented (unless some effects of quantum gravity remedy the situation) by the infinite increase of the energy density. \smallskip \par The goal of this paper is to state that there is actually {\em no\/} reasons to expect that the energy density diverges at the Cauchy horizon in the general case. In particular, we cite a few examples in which the stress-energy tensor of the massless scalar field (as found in the usual way --- through quantization directly in $M$) remains bounded as the horizon is approached. These examples are not of course anywhere near an adequate model of the time machine (for example, the wave equation on the two-dimensional cylinder has no solutions, except for constant, continuous at the Cauchy horizon). However, they can well serve as counterexamples to statements like that cited above. So it seems important to find out whether they indicate only some loop-holes in the FKT approach, which can then be used in the general case, or whether this approach must be fully revised. Our analysis in Section~2 shows that the latter is true --- by several independent reasons one cannot extract any information from expression (\ref{Tren}). In Section~3 some special cases are considered necessary to support statements from Section~2. \section{Analysis} \subsection{Going to the universal covering} Formula (\ref{Cov}), combined with some implicit assumptions serves as a basis for the overall FKT approach since one cannot use (\ref{Had}) in multiply connected spacetimes, where $\sigma$ is not defined. To discuss its validity and to reveal these assumptions let us first state the simple fact that most properties of the Hadamard function, {\em including\/} the validity of (\ref{Had}), depends on the choice of the vacuum appearing in definition (\ref{DefG}). So, formulas like (\ref{Cov}) are meaningless until we specify the vacuua $|0\rangle $ and $|\tilde 0\rangle $ in $M$ and ${\widetilde M}$ respectively. We come thus to the problem of great importance in our consideration --- how, given $|0\rangle $, one could determine corresponding $|\tilde 0\rangle $? The above-mentioned assumptions concern just this problem. They must be something like the following \begin{enumerate} \item For any vacuum on $M$ there exists a vacuum $|\tilde 0\rangle $ on $ \widetilde{M}$ such that (\ref{Cov}) holds. \item The function $\widetilde{G}^{(1)}$ corresponding to $|\tilde 0\rangle $ has the "Hadamard form" (\ref{Had}). \item ${G}^{(1)}$ determines $\widetilde{G}^{(1)}$ uniquely. \end{enumerate}\par The validity of Assumption 1 is almost obvious in the simplest cases (see below), but it was not proven in the general case. (One can meet the references to \cite{Ban} in this connection. Note, however, that the functions $K_C$ which stand there in the analog of our formula (\ref{Cov}), are actually not defined% \footnote{I am grateful to Dr. Parfyonov, who explained to me this issue.} in our case, i.~e.\ when $|\Gamma|=\infty$.) Assumption 2 seems still more arbitrary. The validity of (\ref{Had}) was proven not for {\em any\/} state, but only for some specific class of states (see \cite[Sect.~2c]{Ful}) and there is no reason to believe that our $% |\tilde0\rangle$ belongs just to this class. Assumption 3 is definitely untrue. In the following section we construct as an example a class of vacuua $|\tilde0\rangle_f$ such that (\ref{Cov}) is satisfied for any $f$ while $\widetilde{G}_f$ differ for different $f$. This nonuniqueness is far from harmless. As we argue below it makes, in fact, expression (\ref{Tren}) meaningless. \subsection{The expression for the stress-energy tensor} Expression (\ref{Tren}) is the main result of the FKT approach and (\ref{Cov}) is needed only to justify it. So let us state first that \begin{enumerate} \item (\ref{Tren}) does not follow (or, at least, does not follow immediately) from (\ref{Cov}), since \begin{enumerate} \item To write $\lim D_{\mu \nu }\sum \widetilde G^{(1)}=\sum \lim D_{\mu \nu } \widetilde G^{(1)}$ without a special proof one must be sure that the series $\sum \widetilde G^{(1)}$ and $\sum D_{\mu \nu }\widetilde G^{(1)}$ converge uniformly, while it is clear that they do not (at least as long as (\ref{Had}% ) holds). This nonuniformity manifests itself, in particular, in the fact that, in general, one cannot drop the nonsingular terms in $\widetilde G^{(1)}$. In Subsection~$3.2$ we shall show that the last series in (\ref{Tren}) can diverge {\em off\/} the Cauchy horizon though (\ref{Had},\ref{Cov}) hold and $G^{(1){\rm ren}}$ (and $\langle {\bf T}\rangle^{{\rm ren}}$) are smooth there. \item Even when $|\tilde 0\rangle $ belongs to the above-mentioned class, (% \ref{Had}) is proven not for {\em any\/} $X,\ X^{\prime }$, but only for $% X^{\prime }$ lying in the ''sufficiently small'' neighborhood of $X$. It is necessary, in particular, that $\sigma (X,X^{\prime })$ would be defined uniquely. To provide this in Ref. \cite{Ful}, for example, $X$ and $% X^{\prime }$ are required not to lie respectively near points $\underline{\vphantom{y}x}, \mbox{ and }\underline{y}$ connected by a null geodesic with a point conjugate to $\underline{\vphantom{y}x}$ before $\underline{y}$. To violate this condition for the points $X'$ and $\gamma X'$ it suffices to separate the mouths of the wormhole widely enough and to fill the space between them with the conventional matter \cite[Prop. 4.4.5]{HawEl}. \end{enumerate} \vbox{Thus, we see that (\ref{Tren}) must be regarded as an independent assumption. We can, however, neither use nor check it in view of the aforementioned ambiguity:} \item Like the Hadamard function, $\langle T_{\mu\nu}\rangle^{{\rm ren}}_{\widetilde M}$ depends on which vacuum we choose, while from the FKT standpoint all vacuua $|\tilde0\rangle$ satisfying (\ref{Cov}) are equivalent. This equivalence is of a fundamental nature --- the only physical object is the spacetime $M$, while $\widetilde M$ and $|\tilde0\rangle$ are some auxiliary matters and as long as (\ref{Cov}) holds we cannot apply any extraneous criteria to distinguish among them. So, we have no way of determining what to substitute in (\ref{Tren}) as $\langle T_{\mu\nu}\rangle^{{\rm ren}}_{\widetilde M}$. In Subsection~$3.2$ we shall see that choosing different $|\tilde0\rangle$ (even when $\widetilde M$ is a part of the Minkowski plane) one can make $\langle T_{\mu\nu}\rangle^{{\rm ren}}_{\widetilde M}$ finite or infinite at the horizon at will.\par \end{enumerate} Let me note in passing that there is no point in using (\ref{Tren}) unless we decide that $|\tilde0\rangle$ is among the very "good" and convenient vacuua. For an arbitrary $|\tilde0\rangle$, it is not a bit easier to find $\langle {\bf T}\rangle^{{\rm ren}}_{\widetilde M}$ than $\langle {\bf T}\rangle^{{\rm ren}}_{M\vphantom{\widetilde M}}$. \subsection{Interpretation} Suppose that $\langle\Psi|{\bf T}|\Psi\rangle^{{\rm ren}}_M$ for some $|\Psi\rangle $ does diverge at the Cauchy horizon. Suppose further that it is $\langle {\bf T}\rangle^{{\rm ren}}_M$ that stands in the right side of the Einstein equations (though it is not obvious, see \cite{Bir} for the literature and discussion). Does this really mean that owing to the quantum effects the time machine $M$ cannot be created? I think that the answer is negative. It well may be that $\langle\Phi|{\bf T}|\Phi\rangle^{{\rm ren}}_M$ does not diverge for some other state $ |\Phi\rangle$ (example see in Subsection~$3.3$). Why must we restrict ourselves to the state $|\Psi\rangle$? To prove that the Einstein equations and QFT are incompatible in $M$ one must have proven that the expected stress-energy tensor tends to infinity for {\em any\/} quantum state, or at least for any state satisfying some reasonable physical conditions (say, stability). \section{Examples} Let us find the expectation value of the stress-energy tensor in a few specific cases. We restrict our consideration to the two-dimensional cylinder $M$ obtained from the plane $(\tau ,\chi )$ by identifying $\chi\leadsto \chi +H $ and endowed with the metric \begin{equation} \label{Metr} ds^2=C(-d\tau^2+d\chi^2)=C\,dudv . \end{equation} Here $u\equiv\chi-\tau,\ v\equiv\chi+\tau$; $C$ is a smooth function on $M$. To find in the ordinary way $\langle {\bf T}\rangle $ for the free real scalar field $\phi $ \begin{equation} \label{Weq}\Box \phi =0,\qquad \phi (\chi +H,\tau ) \cases{\phi(\chi,\tau)&for the non-twisted field\cr -\phi(\chi,\tau)&for the twisted field\cr} \end{equation} we must first of all specify the vacuum we consider. That is we must choose a linear space of solutions of (\ref{Weq}) and an "orthonormal" basis \cite {Bir} $U=\{u_n\}$ in it. In particular, this will define the Hadamard function: $$ G^{(1)}(X,X')=\sum_n u_n(X)u^*_n(X')+\mbox{complex conjugate}. $$ A possible choice of $U$ for the non-twisted field is \begin{equation} u_n=|4\pi n|^{-1/2}e^{2\pi i H^{-1}(n\chi -|n|\tau )} \quad n=\pm1,\pm2\dots \label{ConVac} \end{equation} The vacuum $|0\rangle_C$ defined by (\ref{ConVac}) (the "conformal" vacuum) is especially attractive as the expressions for the Hadamard function $G_C^{(1)}$ and for the stress-energy tensor $\langle {\bf T}\rangle _C $ are already obtained (see \cite[the neighborhood of formula (6.211)]{Bir}): \begin{eqnarray} \nonumber\langle T_{ww}\rangle _C^{\rm ren} &=& -\frac{\pi\epsilon}{12H^2}+\frac{1}{24\pi} \left[ \frac{C,_{ww}}{C}-\frac{3}{2}\frac{{C,_w}^2}{C^2}\label{Tcon \right] ,\quad w=u,v\\ &&\strut\\ \nonumber\langle T_{uv}\rangle _C^{\rm ren}&=& \langle T_{vu}\rangle _C^{\rm ren}=-RC/(96\pi). \end{eqnarray} Here $\epsilon=-1/2$ or 1 depending on whether $\phi$ is twisted or untwisted and $R$ is the curvature of $M$. Though the absence of a solution corresponding to $n=0$ in (\ref{ConVac}) may seem artificial, it is, in fact, an inherent feature of $|0\rangle_C$, which is to describe the vacuum of $\phi$ as a massless limit of the "natural" vacuum of a massive field (cf \cite[below (4.220)]{Bir}). One could start, however, from another vacuum for the massive field and arrive at another theory (see below) with the basis $U^{\prime}$ $$ U^{\prime}= U\cup u_0\equiv (2H)^{-1/2}(F\tau+i/F) . $$ Where the real constant $F$ is a free parameter. Choosing different $F\neq0$ we obtain different vacuua $|0\rangle_F$ and Hadamard functions $G^{(1)}_F$. It is easy to see that \begin{equation} \label{GF}G_F^{(1)}=G_C^{(1)}+\frac{F^2}{H}\tau\tau^{\prime}+const. \end{equation} \subsection{Two-dimensional time machines in the conformal vacuum state} As a first example let us consider the Misner spacetime, which is the quadrant $\alpha <0,$ $\beta >0$ of the Minkowski plane $% ds^2=d\alpha d\beta $ with points identified by the rule $(\alpha _{0,}\beta _0)\mapsto (A\alpha _{0,}\beta _0/A)$. The coordinate transformation% $$ u=-W^{-1}\ln |W\alpha |,\quad v=W^{-1}\ln (W\beta ) $$ delivers the isometry between Misner space and $M$ with% $$ C=e^{2W\tau },\qquad H=W^{-1}\ln A . $$ $W$ here is an arbitrary parameter with dimension of mass. Substituting this in (\ref{Tcon}) we immediately find% $$ \langle T_{ww'}\rangle_C^{\rm ren}= -W^2\left(\frac{\epsilon\pi}{12\ln^2A}+\frac{1}{48\pi}\right)\delta_{ww'}. $$ \noindent The metric in coordinates $\alpha,\beta$ is "good" (smooth, nondegenerate) near the Cauchy horizons $\alpha=0$ or $\beta=0$. So, the proper basis of an observer approaching to one of them with a finite acceleration is related to the basis $D\equiv\{\partial_\alpha,\partial_\beta\}$ by a finite Lorentz transformation. Thus the quantities we are to examine are, in fact, the components of $\langle {\bf T}\rangle_C^{\rm ren}$ in the basis $D$, which are $$ \begin{array}{c} \langle T_{\alpha\alpha}\rangle_C^{\rm ren}=T\alpha^{-2},\quad \langle T_{\beta\beta}\rangle_C^{\rm ren}=T\beta^{-2},\quad \langle T_{\alpha\beta}\rangle_C^{\rm ren}= \langle T_{\beta\alpha}\rangle_C^{\rm ren}=0, \\ \strut\\ \displaystyle T\equiv -\left(\frac{\pi\epsilon}{12\ln^2A}+ \frac{1}{48\pi}\right). \end{array} $$\par Now let us use the above simple method to find $\langle {\bf T}\rangle^{\rm ren}$ for two time machines more (see also \cite{Yur}). Consider first the cylinder $S$ obtained from the strip \begin{equation} \begin{array}{c} ds^2=W^{-2}\xi^{-2}(-d\eta^2+d\xi^2)=\xi^{-2}d\alpha d\beta, \\ \strut \\ \mbox{where }\alpha\equiv(\xi-\eta)/W,\ \beta\equiv(\xi+\eta)/W;\quad \eta\in(-\infty,\infty),\;\xi\in[1,A]. \label{StMod \end{array} \end{equation} by gluing points $\eta=\eta_0,\ \xi=1$ with the points $\eta=A\eta_0,\ \xi=A$. This spacetime was considered in detail in \cite{Fro} where it was called the "standard model". A simple investigation shows that the Cauchy horizons $\alpha=0$ and $\beta=0$ divide $S$ into three regions. Causality holds in the "inner" region $\widetilde S:\ \alpha, \beta>0$ and violates in $I^\pm(\widetilde S)$. Introducing new coordinates $u,\;v$: $$ u\equiv W^{-1}\ln \alpha,\quad v\equiv W^{-1}\ln \beta $$ we find that $\widetilde S$ like the Misner space% \footnote{In spite of their apparent similarity these spaces are significantly distinct. For example, the Misner spacetime is geodesically incomplete \cite{HawEl}, and the standard model is not \cite{Me}. This may be of importance if one would like to separate $X$ and $X'$ "widely enough" (see item 1b in the previous section).} is isometric to $M$. This time $$ C=\cosh^{-2}W\tau,\quad H=W^{-1}\ln A, $$ which yields \begin{equation} \begin{array}{c} \langle T_{\alpha\alpha}\rangle_C^{\rm ren}=T\alpha^{-2},\quad \langle T_{\beta\beta}\rangle_C^{\rm ren}=T\beta^{-2},\\ \strut\\ \langle T_{\alpha\beta}\rangle_C^{\rm ren}= \langle T_{\beta\alpha}\rangle_C^{\rm ren}=(1/12\pi)(\alpha+\beta)^{-2}. \label{TFro} \end{array} \end{equation}\par Consider lastly the spacetime obtained by changing $\xi^{-2}\to\eta^{-2}$ in (\ref{StMod}). This spacetime is similar to the standard model, but has a somewhat more curios causal structure --- there are two causally nonconnected regions separated by the time machine. $\langle {\bf T}\rangle_C^{\rm ren}$ differs from that in (\ref{TFro}) by the off-diagonal (bounded) terms $$ \langle T_{\alpha\beta}\rangle_C^{\rm ren}= \langle T_{\beta\alpha}\rangle_C^{\rm ren}=-(1/12\pi)(\alpha-\beta)^{-2}. $$\par So, we see that in all three cases the vacuum energy density (associated with {\em some} vacuum states) does grow infinitely as one approaches to the Cauchy horizon. A few comments are necessary, however: \begin{enumerate} \item The divergence in discussion is not at all something peculiar to the time machine: the passage to the limit $A\to\infty$ shows that precisely the same divergence (with $T=-1/(48\pi)$) takes place in $\widetilde M$ though (in the case of Misner space) $\widetilde M$ is merely a part of the Minkowski plane. This suggests that for the time machine too, the divergence of the stress-energy tensor is a consequence not of its causal or topological structure but rather of the unfortunate choice of the quantum state. \item The twisted field at $A=e^{\sqrt{2}\pi}$ has the bounded $\langle {\bf T}\rangle_C^{\rm ren}$ (cf. \cite{Sush}). \item Let us consider nonvacuum states now (see Subsection~$2.3$). The first example is a two-particle state $|1_n1_{-n}\rangle$ with the particles corresponding to the $n$-th and $-n$-th modes of (\ref{ConVac}). $\langle1_{-n}1_n| {\bf T}|1_n1_{-n}\rangle^{\rm ren}$ is readily found using \cite[eq.~(2.44)]{Bir}: $$ \langle1_{-n}1_n| T_{\gamma\gamma}|1_n1_{-n}\rangle^{\rm ren}=T'\gamma^{-2},\quad \langle1_{-n}1_n| T_{\alpha\beta}|1_n1_{-n}\rangle^{\rm ren} =\langle T_{\alpha\beta}\rangle^{\rm ren}_C $$ with $T'\equiv T+2\pi nH^{-2},$ and $ \gamma\equiv\alpha,\beta$. Thus we see that there {\em are\/} states with the bounded energy density of the untwisted field.\par Another yet example is the equilibrium state at a nonzero temperature $|t\rangle$. Expression (4.27) of \cite{Bir} gives $$ \langle t| T_{ww}| t\rangle^{\rm ren} =\langle T_{ww}\rangle^{\rm ren}_C+\frac{\pi}{2H^2}\sum_{m=1}^\infty \sinh^{-2}\frac{\pi m}{k_BtH}. $$ So, for any $H$ there exists such temperature $t$ that $\langle t| T_{\gamma\gamma}| t\rangle^{\rm ren}$ does not diverge at the horizon. \end{enumerate} \subsection{Another vacuum} The conformal vanuum is not suited for verifying or exemplifying most of statements made in Section 2., since the Hadamard function does not exist in this state. So consider now the new vacuum $|0\rangle_f$ on the plane $(\tau,\chi)$ defined by the modes \begin{equation} u'_p\equiv\cases{ \displaystyle\frac{1}{ 2\sqrt{\pi\omega} } e^{ip\chi-i\omega\tau},& $\omega\geq\delta$\cr \strut&\cr \displaystyle\frac{1}{ 2\sqrt{\pi} } e^{ip\chi} (f^{-1}\cos\omega\tau-i\omega^{-1}f\sin\omega\tau),& $\omega<\delta$.\cr} \label{AnVac} \end{equation} where $\omega\equiv|p|$, $\delta$ is an arbitrary positive constant: $\delta<1$ and $f$ is an arbitrary smooth positive function: $f(\omega\geq \delta)=\sqrt\omega$. The modes (\ref{AnVac}) are obtained >from that defining the conformal vacuum on the plane by a Bogolubov transformation of the low-frequency modes so as to avoid the infrared divergence without affecting the ultraviolet behavior of $\langle {\bf T}\rangle$. The asymptotic form of ${\widetilde G}^{(1)}_f$ does not depend on $f$: \begin{equation} \forall f\qquad \widetilde G^{(1)}_f=-1/(2\pi)\ln|\Delta u\Delta v|+\mbox{smooth, bounded function} . \label{GAs} \end{equation} If we retain only the first term, we obtain (in the flat case) $$ \lim_{X^{\prime}\to X} D_{\alpha\alpha}{\widetilde G}^{(1)}(X,\gamma^nX')= \ln^2(A)\, \alpha^{-2}A^{-n}n^{-2}. $$ So, the last series in (\ref{Tren}) diverges not only at the horizon, but everywhere on $M$ (cf. Subsection $2.2$).\par $\langle{\bf T}\rangle_f$ can be found from (\ref{GF}) (see \cite[Sect.~$6.4$]{Bir}). For any $C$ we have: $$ \langle T_{ww}\rangle_f^{\rm ren}= \frac{1}{8\pi}\left[ -\frac{\;W^2}{6}+ \int^\delta_0 (f^{-2}\omega^2 +f^2-\omega) \,d\omega\right],\quad \langle T_{uv}\rangle_f^{\rm ren}= \langle T_{uv}\rangle_C^{\rm ren}. $$ Having taken an appropriate $f(\omega)$ one can make $\langle T_{\alpha\alpha}\rangle^{\rm ren}_f$ infinite or zero at the horizon, as we have stated in Subsection $2.2$.\par To illustrate some more statements from Section~2.\ let us first find $G^\Sigma$. To this end note that it has the form \begin{equation} G^\Sigma=\sum_n\int_{-\infty}^{\infty}h(p)e^{inHp}\,dp + c.\ c. \label{GSum \end{equation} with $$ h\equiv\cases{ \displaystyle\frac{1}{ 4\pi\omega } e^{ip\Delta\chi-i\omega\Delta\tau},& $\omega\geq1/2$\cr \strut&\cr \displaystyle\frac{1}{ 4\pi } e^{ip\Delta\chi} (f^{-2}\cos\omega\tau\cos\omega\tau'+ \omega^{-2}f^2\sin\omega\tau\sin\omega\tau') ,& $\omega<1/2$.\cr} $$ The function $h(p)$ can be written as a sum: $h=(h-h_0)+h_0$, where $$ h_0\equiv\frac{1}{4\pi\sqrt{1+p^2}}\;e^{ip\Delta\chi} \; e^{-i\sqrt{1+p^2}\Delta\tau}. $$ The first summand is a smooth function falling off at infinity like $p^{-2}$ and the second summand ($h_0$) is a holomorphic (but for $p=\pm i$) function admitting the following estimate: $$ |h_0|\leq C|x|^{-1/2}\;e^{|(\Delta\chi-\Delta\tau)y|}. $$ Hence \cite{Fed} we can apply the Poisson formula to (\ref{GSum}) and obtain: $$ G^\Sigma=2\pi H^{-1}\sum_n h(2\pi H^{-1}n) +c.\ c. $$ We see thus that $G^\Sigma$ is indeed the Hadamard function and it corresponds to the vacuum $|0\rangle_F$ with $F=f(0)$.\par {\em Remark 1.\/} This does not mean, however, that $G^\Sigma$ will be a Hadamard function of some reasonable state for {\em any\/} $\widetilde G^{(1)}$. One can easily construct, for example, such a vacuum that $\widetilde G^{(1)}(\chi,\chi')$ will not be invariant under translations $\ \chi,\chi'\mapsto\chi+H,\chi'+H$ and $G^\Sigma(\chi,\chi')$, as a consequence, will not even be symmetric.\par {\em Remark 2.\/} For all $G^{(1)}_f$ with the same $f(0)$ the Hadamard functions $G^\Sigma$ are the same. This proves our statement from Subsection~$2.1$. To find $\langle {\bf T}\rangle_F$ note that it differs from $\langle {\bf T}\rangle_C$ only by the term arising from the second summand in (\ref{GF}) (cf.~\cite[ eqs. (4.20), (6.136)]{Bir}): $$ \Delta\langle T_{ww'}\rangle^{\rm ren}=\frac{F^2}{2H} (\tau,_w\tau,_{w'}- 1/2\,\eta_{ww'}\eta^{\lambda\delta}\tau,_\lambda\tau,_\delta). $$ That is \begin{eqnarray*} \langle T_{ww}\rangle_F^{\rm ren}&=& \displaystyle{\frac{F^2}{8H\vphantom{\ln^2A}}- W^2\left(\frac{\epsilon\pi}{12\ln^2A}+ \frac{1}{48\pi}\right)},\\ &&\strut\\ \langle T_{uv}\rangle_F^{\rm ren}&= & \langle T_{vu}\rangle_F^{\rm ren}= \langle T_{uv}\rangle_C^{\rm ren} . \end{eqnarray*} So, for all three time machines considered here there exists a vacuum, such that the expectation value of the stress-energy tensor is bounded in the causal region. \section{Conclusion} Thus, we have seen that one cannot obtain any information about the energy density near the Cauchy horizon employing the FKT approach. In the absence of any other general approach this means that all we have is a few simple examples. In some of them the energy density diverges there and in some do not. So, the time machine perhaps is stable and perhaps is not. This seems to be the most strong assertion we can make.
\subsection{The case of zero mixing} In this case $A_t=0$, and the expression for $G$ in (\ref{G}) simplifies a lot. Using (\ref{redefinicion}) and (\ref{betasns}) one can, to one loop, expand $\lambda$ as \begin{eqnarray} \label{exp} \lambda(\overline{m}_{t})&=& \lambda(M)-\beta_{\lambda}(M)\log\frac{M^2}{\overline{m}_{t}^2} \nonumber \\ &=& \frac{1}{4}(g^2+g'^2)+\frac{3}{8\pi^2}h_t^2(h_t^2-\lambda) \log\frac{M^2}{\overline{m}_{t}^2} \end{eqnarray} while ${\rm d}^2G/{\rm d}x^2$ is found to give \begin{equation} \label{Gseg} \frac{{\rm d}^2G}{{\rm d}x^2} = \frac{3}{8\pi^2} h_t^2(h_t^2-\lambda)\log\frac{M^2+\overline{m}_{t}^2}{M^2} -\frac{3}{16\pi^2}h_t^4\frac{M_Z^2}{M^2+\overline{m}_{t}^2} \end{equation} where all couplings are considered at the scale $M$ (remember that the $G$-term was frozen at that scale) and we have neglected ${\cal O}(g^4)$ terms. Putting (\ref{exp}) and (\ref{Gseg}) together, the effective coupling (\ref{lambdaeffec}) can be written as \begin{equation} \label{efflambda} \lambda_{\rm eff}(\overline{m}_{t})=\frac{1}{4}(g^2+g'^2) +\frac{3}{8\pi^2}h_t^2(h_t^2-\lambda_{\rm eff}) \log\frac{M^2+\overline{m}_{t}^2}{\overline{m}_{t}^2} -\frac{3}{16\pi^2}h_t^4\frac{M_Z^2}{M^2+\overline{m}_{t}^2} \end{equation} where all couplings on the right-hand side are considered at the scale $M$. Comparison of (\ref{exp}) with (\ref{efflambda}) shows that the effect of including the $G$-term in (\ref{potencialsub}) is, apart from the last (mixed Yukawa--gauge) term in (\ref{efflambda}) which is numerically unimportant, to resum the series $$ \sum_{n=1}^{\infty}\frac{(-1)^{n+1}}{n} \left(\frac{\overline{m}_{t}^{2n}}{M^{2n}}\right)$$ from the expansion parameter $\log(M^2/\overline{m}_{t}^2)$ to the physical one $\log((M^2+\overline{m}_{t}^2)/\overline{m}_{t}^2)$, which goes to zero in the supersymmetric limit $M\rightarrow 0$, and makes the radiative corrections vanish in that limit. In other words the resummation from the $G$-function allows us to replace the expansion (\ref{exp}) by the expansion \begin{equation} \label{expMs} \lambda_{\rm eff}(\overline{m}_{t})= \lambda_{\rm eff}(M_S)-\beta_{\lambda_{eff}}(M_S)t \end{equation} where the expansion parameter is \begin{equation} \label{t} t=\log\frac{M_S^2}{\overline{m}_{t}^2} \end{equation} and \begin{equation} \label{MS} M_S^2=M^2+\overline{m}_{t}^2 . \end{equation} \subsection{The case $A_t\neq 0$} Using again (\ref{redefinicion}) and (\ref{betasns}) one can write \begin{eqnarray} \label{expmix} \lambda(\overline{m}_{t}) & = & \lambda(M)- \beta_{\lambda}(M)\log\frac{M^2}{\overline{m}_{t}^2}\nonumber \\ &=& \frac{1}{4}(g^2+g'^2)+\frac{3}{8\pi^2}h_t^2(h_t^2-\lambda) \log\frac{M^2}{\overline{m}_{t}^2} \\ &+ & \frac{3}{8\pi^2} h_t^2 \left[\left(h_t^2-\frac{1}{8}(g^2+g'^2)\right)\frac{A_t^2}{M^2} -\frac{1}{12}h_t^2\frac{A_t^4}{M^4}\right] \nonumber \end{eqnarray} In the absence of the $G$-contribution in (\ref{lambdaeffec}), Eq.~(\ref{expmix}) gives the usual one-loop leading-log contribution to the Higgs mass. Introducing now the $G$-function, one can write (\ref{lambdaeffec}) as: \begin{equation} \label{efflamix} \lambda_{\rm eff}(\overline{m}_{t}) = \frac{1}{4}(g^2+g'^2)+\frac{3}{8\pi^2}h_t^2 (h_t^2-\lambda_{\rm eff}) \log\frac{M_S^2}{\overline{m}_{t}^2} + \Delta_{\rm th}\lambda_{\rm eff} , \end{equation} where the `threshold' contribution to the coupling is given by \begin{eqnarray} \label{umbral} \Delta_{\rm th}\lambda_{\rm eff} & = & \frac{3}{64 v^4\pi^2}\left\{\left(2m_t^2-\frac{1}{2}M_Z^2+m_t A_t\right)^2 \log\left(1+\frac{m_tA_t-\frac{1}{4}M_Z^2}{M_S^2}\right) \right. \nonumber \\ &+ & \left(2m_t^2-\frac{1}{2}M_Z^2-m_t A_t\right)^2 \log\left(1-\frac{m_tA_t+\frac{1}{4}M_Z^2}{M_S^2}\right) \\ &- & m_t A_t \left\{ \left(M_S^2-\frac{1}{4}M_Z^2+m_t A_t \right) \left[\log\left(1+ \frac{m_t A_t-\frac{1}{4}M_Z^2}{M_S^2}\right)-1\right]\right. \nonumber \\ & - & \left.\left. \left(M_S^2-\frac{1}{4}M_Z^2-m_t A_t \right) \left[\log\left(1- \frac{m_t A_t+\frac{1}{4}M_Z^2}{M_S^2}\right)-1\right]\right\} \right\} \nonumber \end{eqnarray} and it can be easily checked that \begin{equation} \label{expumbral} \Delta_{\rm th}\lambda_{\rm eff}= \frac{3}{8\pi^2} h_t^2\left[\left(h_t^2-\frac{1}{8}(g^2+g'^2)\right) \frac{A_t^2}{M_S^2} -\frac{1}{12}h_t^2\left(\frac{A_t^2}{M_S^2}\right)^2\right] +\cdots \end{equation} where the ellipsis denotes the contribution from higher-dimensional terms in $m_t^2/M_S^2$ and $A_t^2/M_S^2$. We can see that the effect of taking into account the $G$-function in (\ref{potencialsub}) is twofold: \begin{itemize} \item It makes a resummation from $\log(M^2/\overline{m}_{t}^2)$ to the physical expansion parameter $\log(M_S^2/\overline{m}_{t}^2)$ in the one-loop radiative correction, as well as resummation from $A_t^2/M^2$ to $A_t^2/M_S^2$ in the threshold term of (\ref{expmix}). \item It generalizes the threshold correction of Eq.~(\ref{expmix}) to include higher order effects in powers of $A_t^2/M_S^2$ in Eq.~(\ref{umbral}). \end{itemize} Therefore the net effect of the $G$-function is to change the expansion (\ref{expmix}) into \begin{equation} \label{expMsmix} \lambda_{\rm eff}(\overline{m}_{t})=\lambda_{\rm eff}(M_S)- \beta_{\lambda_{\rm eff}}(M_S)t \end{equation} where \begin{equation} \label{inicial} \lambda_{\rm eff}(M_S)= \frac{1}{4}(g^2+g'^2)+\Delta_{\rm th}\lambda_{\rm eff} . \end{equation} \subsection{Two-loop leading-log expansion} As we have seen in Ref.~\cite{CEQW} one can obtain the two-loop leading-log correction by expanding the parameter $\lambda$ to order $(\log(M^2/\overline{m}_{t}^2))^2$, as \begin{equation} \label{expmix2} \lambda(\overline{m}_{t})=\lambda(M)-\beta_{\lambda}(M)\log\frac{M^2}{\overline{m}_{t}^2} +\frac{1}{2}\beta'_{\lambda}\left(\log\frac{M^2}{\overline{m}_{t}^2}\right)^2, \end{equation} where, since we are now considering two-loop corrections, we have evaluated the quartic coupling at the on-shell top quark mass scale and the prime denotes derivative with respect to $\log(Q^2)$. In the previous subsection we have seen that the effect of including the $G$-function in the effective potential is to resum $M^2$ to $M_S^2$ in the first two terms of (\ref{expmix2}) and replace the threshold correction in the first one by $\Delta_{\rm th}\lambda_{\rm eff}$ in (\ref{umbral}) (in Ref.~\cite{CEQW}, this resummation effect was assumed to be true to assure a proper behaviour of the radiative corrections for $M = 0$). To prove resummation in the last term of (\ref{expmix2}) one would need to use the whole two-loop effective potential in the MSSM. In the absence of such a calculation we shall assume that this happens, since we have already proved that $t$ in (\ref{t}) is the physical expansion parameter in the one-loop calculation, and in addition the numerical relevance of the resummation in the two-loop corrections is expected to be tiny. We shall therefore consider as a starting point \begin{eqnarray} \label{efflamix2} \lambda_{\rm eff}(\overline{m}_{t})& = & \lambda_{\rm eff}(M_S)-\beta_{\lambda_{\rm eff}}(M_S)t +\frac{1}{2}\beta'_{\lambda_{\rm eff}}(\overline{m}_{t})t^2 +\cdots \nonumber \\ & = & \lambda_{\rm eff}(M_S)-\beta_{\lambda_{\rm eff}}(\overline{m}_{t})t -\frac{1}{2}\beta'_{\lambda_{\rm eff}}(\overline{m}_{t})t^2+\cdots \end{eqnarray} where $\lambda_{\rm eff}(M_S)$ is given in (\ref{inicial}) and $t$ and $M_S$ are obtained from Eqs.~(\ref{t}) and (\ref{MS}). Defining $\beta_{\lambda_{\rm eff}} = a_{\lambda_{\rm eff}} \lambda_{\rm eff} + b_{\lambda_{\rm eff}}$, it follows that \begin{equation} \lambda_{\rm eff}(\overline{m}_{t}) = \lambda_{\rm eff}(M_S) \left( 1 - a_{\lambda_{\rm eff}}(\overline{m}_{t}) \; t \right) - b_{\lambda_{\rm eff}}(\overline{m}_{t}) \; t \; \left( 1 - a_{\lambda_{\rm eff}}(\overline{m}_{t}) \; t \right) -\frac{1}{2}\beta'_{\lambda_{\rm eff}}(\overline{m}_{t})t^2. \end{equation} {}From the renormalization group equations of the quartic coupling, it follows that the term $1 - a_{\lambda_{\rm eff}}(\overline{m}_{t}) \;t = \xi^{-4}(\overline{m}_{t})$, where $\xi$, the Higgs field anomalous dimension, is: \begin{equation} \label{anomdim} \xi(\overline{m}_{t})=1+\frac{3}{32\pi^2}h_t^2(\overline{m}_{t})\;t. \end{equation} Recalling that $v^2(M_S) = v^2(\overline{m}_{t}) \; \xi^{-2}(\overline{m}_{t})$, from Eq.~(\ref{masahiggs}) we get \begin{equation} \label{decmasa} m_h^2(\overline{m}_{t}) = m_h^2(M_S)\;\xi^{-2}(\overline{m}_{t})+ \Delta_{\rm rad\; }m_h^2(\overline{m}_{t}). \end{equation} In the above, \begin{equation} \label{masafroz} m_h^2(M_S) = 2 \lambda_{\rm eff}(M_S)\;v^2(M_S), \end{equation} $\lambda_{\rm eff}(M_S)$ is given in Eq.~(\ref{inicial}) with all couplings and masses evaluated at the scale $M_S$, and $\Delta_{\rm rad\; }m_h^2(\overline{m}_{t})$ is given by \begin{eqnarray} \label{radmas} \Delta_{\rm rad\; }m_h^2(\overline{m}_{t})& = & 2v^2(\overline{m}_{t})\left[ - b_{\lambda_{\rm eff}}(\overline{m}_{t}) \; t \; \left( 1 - a_{\lambda_{\rm eff}}(\overline{m}_{t}) \; t \right) -\frac{1}{2}\beta'_{\lambda_{\rm eff}}(\overline{m}_{t})t^2 \right] \nonumber \\ & = & \frac{3}{4\pi^2}\frac{\overline{m}_{t}^4}{v^2(\overline{m}_{t})}\;t \left[1+\frac{1}{16\pi^2}\left(\frac{3}{2}h_t^2-32\pi\alpha_3\right)t \right] \end{eqnarray} where all couplings in (\ref{radmas}) are evaluated at the scale $Q^2=\overline{m}_{t}^2$, $\alpha_3$ is the strong gauge coupling, and we have used $\beta_{h_t}$ in the evaluation of (\ref{radmas})~\cite{CEQW}. Observe that, if we replace $\Delta_{\rm th}\lambda_{\rm eff}$ by its expansion in powers of $A_t/M_S$, Eq.~(\ref{expumbral}), we neglect the small terms depending on the weak gauge couplings, and we re-express the values of the couplings at $M_S$ by their expressions at $\overline{m}_{t}$ using the appropriate $\gamma$- and $\beta$-functions, we obtain \begin{eqnarray} m_h^2& = & M_Z^2\left( 1-\frac{3}{8\pi^2}\frac{\overline{m}_{t}^2} {v^2}\ t\right) \nonumber \\ \label{mhsm} & + & \frac{3}{4\pi^2}\frac{\overline{m}_{t}^4}{v^2}\left[ \frac{1}{2}\tilde{X}_t + t +\frac{1}{16\pi^2}\left(\frac{3}{2}\frac{\overline{m}_{t}^2}{v^2}-32\pi\alpha_3 \right)\left(\tilde{X}_t t+t^2\right) \right] \end{eqnarray} where \begin{eqnarray} \label{stopmix} \tilde{X}_{t} & =& \frac{2 A_t^2}{M_{S}^2} \left(1 - \frac{A_t^2}{12 M_{S}^2} \right). \end{eqnarray} Equation (\ref{mhsm}) is equivalent to Eq.~(9) of Ref.~\cite{CEQW} in the large $\tan\beta$ regime. Although we have used, as simplifying hypothesis, the case where $\tan\beta\gg 1$ and $\mu = 0$, all the results are also valid for the case $m_A\sim M_S$ and any value of $\tan\beta$ and $\mu$. The only change in the final results is that \begin{eqnarray} \label{relacion} (g^2+g'^2) & \longrightarrow & (g^2+g'^2)\cos^2 2\beta \nonumber \\ A_t & \longrightarrow & A_t-\mu\cot\beta \end{eqnarray} in $\lambda(M)$ and $\lambda_{\rm eff}(M_S)$. Finally, we want to conclude this section with a comment about the physical interpretation of the decomposition (\ref{decmasa}). The term $m_h^2(M_S)$ comes from the scale independent part of the MSSM effective potential frozen at the scale $M_S$, where we have already subtracted the contribution evolving with $\log(M_S^2/\overline{m}_{t}^2)$. Since this term is scale independent, it is evolved to the scale $\overline{m}_{t}$ with the corresponding power of the anomalous dimension of the Higgs field. On the contrary, the term $\Delta_{\rm rad\; }m_h^2$, which arises from renormalizable terms after the stops are decoupled at the high scale, is computed at the low scale $\overline{m}_{t}$. Had we considered the whole MSSM effective potential at the scale $\overline{m}_{t}$~\cite{ERZ}, we would have been neglecting the stop decoupling at the scale $M_S$. This is shown in Fig.~1 (dotted lines) where we plot $m_h$ as a function of $M_S$ for a pole top-quark mass $M_t=175$ GeV, vanishing mixing, $A_t=\mu=0$, and large ($\tan\beta=15$) and small [infrared (IR) fixed point solution: $\sin\beta\sim (200\; {\rm GeV}/M_t)$] values of $\tan\beta$. Had we evolved the Higgs mass obtained from the whole MSSM effective potential (including the logarithmic terms) at $M_S$, to the scale $\overline{m}_{t}$ with the anomalous dimension factor $\xi^{-2}(\overline{m}_{t})$, we would have made an error associated with the non-exact scale invariance of the effective potential in the one-loop approximation, as was observed in~\cite{CEQR}. This is shown in the dashed lines of Fig.~1 and was also noticed in Ref. \cite{CEQW}. In fact the latter procedure would lead to an expression of $\Delta_{\rm rad\; }m_h^2$, where the second term inside the square brackets has an extra factor of 2 [see Eq.~(\ref{mhsm})]. The solid lines in Fig.~1 show the corresponding value for the Higgs mass, while considering Eq.~(\ref{decmasa}) with the whole expression for $\Delta_{\rm th} \lambda_{\lambda_{\rm eff}}$ from Eq.~(\ref{umbral}). The dependence of (\ref{decmasa}) on the mixing is shown in Fig.~2 (solid lines) where we plot $m_h$ as a function of $A_t$, for $\mu=0$, the same values of $\tan\beta$ as in Fig.~1 and $M_S=1$~TeV. For comparison we also plot (dashed lines) the corresponding mass, using the approximate expression for the threshold contribution to the Higgs quartic couplings, Eq.~(\ref{mhsm}). We see that this approximation, which was used in~\cite{CEQW}, is very good up to the maximum of the curve, which means that the absolute upper bound on the mass of the lightest Higgs boson previously obtained remains unchanged for any value of the mixing. Of course for very large values of the mixing there is a departure between the two curves. We have deliberately omitted from Figs.~1 and 2 the small D-term contributions in Eq.~(\ref{umbral}) to compare with previous results which did not make use of them \cite{ERZ,CEQW}. They will be included in the general analysis of section 3. \setcounter{equation}{0{The general case} Having understood the simplified case explained in section 2, we can proceed with the general case. Let us first focus on the behaviour of the renormalized Higgs quartic couplings for low values of the CP-odd Higgs mass and values of the mass parameter $m_Q$ different from $m_U$. In this case, the effective potential at scales above ${\rm Max}(m_Q,m_U,m_D)$ is obtained through the general expression, Eq.~(\ref{potgeneral}). Expanding the effective potential in powers of the Higgs fields, it is easy to identify the form of the effective quartic couplings at the scale $m_Q$, $\lambda_i(m_Q)$, in a way completely analogous to what we have done for the case $m_Q = m_U$ in section 2. Assuming for definiteness that $m_Q > m_U > m_D$, \begin{equation} \lambda_i(m_Q) = \lambda_i(m_U) + \beta_{i}^{QU}(m_U) \; \widetilde{t}_{QU} + \left(\beta_{i}^{QU}\right)' \; \frac{\;\widetilde{t}_{QU}^{\; 2}}{2} \end{equation} [see e.g. the second equality in Eq.~(\ref{efflamix2})] where $\widetilde{t}_{QU} = \log(m_Q^2/m_U^2)$, $\beta_i$ denotes the $\beta$-function of the Higgs quartic couplings, $\beta_i = d \lambda_i/d\log(Q^2)$ and the superscript $QU$ denotes its expression at energy scales between $m_Q$ and $m_U$. In general, we can write, \begin{equation} \beta_i = a_i \lambda_i + b_i, \;\;\;\;\;\;\;\;\; \beta_i' \simeq a_i b_i + b'_i , \end{equation} where we have only kept the dominant, Yukawa-coupling-dependent contributions to $\beta'_i$. We omit the scale dependence of $\beta_i'$ since it is a higher-order effect. Hence, \begin{equation} \lambda_i(m_U) = \lambda_i(m_Q) \left( 1 - a_i \; \widetilde{t}_{QU} \right) - b^{QU}_i(m_U) \; \widetilde{t}_{QU} \left( 1 - a_i \; \widetilde{t}_{QU} \right) - \left(\beta_{i}^{QU}\right)' \; \frac{\;\widetilde{t}_{QU}^{\; 2}}{2}. \end{equation} The coefficients $a_i$ are linear combinations of the anomalous dimensions of the Higgs fields $H_i$, which are independent of the squark fields. The same procedure as used above may be used to connect the value of the quartic couplings at $m_D$ with their values at $m_U$, and finally the values at $m_D$ with their values at $\overline{m}_{t}$. Keeping only the dominant terms, we obtain \begin{eqnarray} \lambda_i(\overline{m}_{t}) & = &\lambda_i(m_Q) \left( 1 - a_i(\overline{m}_{t}) \; \widetilde{t}_Q \right) - \left(\beta_{i}^{QU}\right)' \frac{\;\widetilde{t}_{QU}}{2} \left(\widetilde{t}_{Q} + \widetilde{t}_{U}\right) \nonumber\\ &-& b^{QU}_i(\overline{m}_{t}) \; \widetilde{t}_{QU} \; \left[ 1 - a_i(\overline{m}_{t}) \left(\widetilde{t}_{Q} + \; \widetilde{t}_{U} \right) \right] \nonumber\\ & - & b^{UD}_i(\overline{m}_{t}) \; \widetilde{t}_{UD} \left[ 1 - a_i(\overline{m}_{t}) \; \left(\widetilde{t}_{U} + \widetilde{t}_D \right) \right] - \left(\beta_{i}^{UD}\right)' \; \frac{\;\widetilde{t}_{UD}}{2} \left(\widetilde{t}_{U} + \widetilde{t}_D \right) \nonumber\\ & - & b_i(\overline{m}_{t}) \; \widetilde{t}_{D} \left( 1 - a_i (\overline{m}_{t})\; \widetilde{t}_{D} \right) -\beta_{i}' \; \frac{\;\widetilde{t}_{D}^{\; 2}}{2}, \label{lambimt} \end{eqnarray} where $\beta_i$ without any superscript denotes the $\beta$-function values once the two stops and the two sbottoms are decoupled, the superscript $XY$, with $X,Y = Q,U,D$, denotes the functional form of the $\beta$-functions at scales between $m_{X}$ and $m_{Y}$, $\widetilde{t}_{XY} = \widetilde{t}_X - \widetilde{t}_Y$ and $\widetilde{t}_X = \log(m_X^2/\overline{m}_{t}^2)$. Similar expressions are obtained for a different hierarchy of the squark mass parameters, with the only difference that, in the case $m_U > m_Q$ and/or $m_D > m_Q$, stops and sbottoms should be decoupled at different scales. For simplicity of presentation we shall first discuss the result for the Higgs mass matrix elements in the case under study and we shall present below the result in the most general case. The contribution of the quartic couplings to the Higgs mass matrix elements is then given by \begin{eqnarray} M^2_{12} &=& 2 v^2 [\sin \beta \cos \beta (\lambda_3 + \lambda_4) + \lambda_6 \cos^2 \beta + \lambda_7 \sin^2 \beta ] - m_A^2 \sin\beta \cos\beta \nonumber\\ M^2_{11} &=& 2 v^2 [\lambda_1 \cos^2 \beta + 2 \lambda_6 \cos \beta \sin \beta + \lambda_5 \sin^2 \beta] + m_A^2 \sin^2\beta \label{mijl}\\ M^2_{22} &=& 2 v^2 [\lambda_2 \sin^2 \beta +2 \lambda_7 \cos \beta \sin \beta + \lambda_5 \cos^2 \beta] + m_A^2 \cos^2\beta, \nonumber \end{eqnarray} where all terms should be evaluated at the same scale and we have also included the dependence on the CP-odd Higgs mass~\cite{HH}. In the above, $v_1 = v \cos\beta$ and $v_2 = v \sin\beta$ are the $H_1$ and $H_2$ vacuum expectation values, respectively. Equation (\ref{lambimt}) has a clear interpretation: The factor $\lambda_i(m_Q)$ in the first term contains the tree level terms and all finite contributions to the quartic couplings, arising from the existence of squark mixing and the fact that $m_D \neq m_Q \neq m_U$. The factor involving $a_i$ in the first term contains exactly the terms necessary to rescale the Higgs mass matrix elements by the appropriate anomalous dimension factors from the scale $m_Q$ to the scale $\overline{m}_{t}$, together with the ones necessary to re-express the vacuum expectation values of the Higgs fields appearing in Eq.~(\ref{mijl}) at the scale $\overline{m}_{t}$, in terms of their values at the scale $m_Q$ (for the complete expression of the $\beta$-functions of the Higgs quartic couplings, see for example Ref.~\cite{HH}). For instance, singling out this contribution of the $\lambda_2$ coupling to the matrix element $M_{22}^2$, which we shall denote by $K_{22}$, we obtain \begin{eqnarray} K_{22}(\overline{m}_{t}) &=&2 \; \lambda_2(m_Q) \left[ 1 - a_2 \log\left( \frac{m_Q^2}{\overline{m}_{t}^2} \right) \right] \; v_2^2(\overline{m}_{t}) \nonumber\\ &=& 2 \; \lambda_2(m_Q) \; \xi_2^{-4}(\overline{m}_{t}) \; v_2^2(\overline{m}_{t}) \\ & & \nonumber \\ &=& 2 \; \lambda_2(m_Q) \; v_2^2(m_Q) \; \xi_2^{-2}(\overline{m}_{t})\nonumber, \end{eqnarray} where $\xi_2$ is the anomalous dimension factor of the Higgs fields $H_2$. The contribution of the other couplings to the Higgs mass matrix elements present similar properties. Thus, both the tree level term and all finite terms leading to the non-trivial matching of the quartic couplings at the scale $m_Q$ may be treated in the same way as the terms proceeding from the higher-dimensional operator contributions to the Higgs mass matrix elements. That is, they may be frozen at the scale $m_Q$ and rescaled with the appropriate anomalous dimension factors to obtain their expressions at low energies. This generalizes the result obtained in section 2 for the case of degenerate squark masses, $m_Q = m_U$. The following terms in Eq.~(\ref{lambimt}) are the ones which would be obtained even in the presence of trivial matching conditions for the quartic Higgs couplings and, as has been clearly explained in the previous section, are associated with the scale-dependent contributions to the effective potential. The dominant leading-log contribution to the quartic couplings proceeds from the terms in the $\beta$-function proportional to the fourth power of the top and bottom quark Yukawa couplings, which are given by \begin{equation} \lambda_i(\overline{m}_{t}) - \lambda_i(m_Q) = - \frac{b_i^Y}{2} \left( \log\left(\frac{m_Q^2}{\overline{m}_{t}^2}\right) + \log\left(\frac{m_X^2}{\overline{m}_{t}^2}\right) \right), \label{llog} \end{equation} where $b_1^Y = -3 h_b^4/8 \pi^2$ and $b_2^Y = -3 h_t^4/8 \pi^2$, $h_b$ and $h_t$ are the bottom and top quark Yukawa couplings, and $X = D, U$ for $i = 1,2$, respectively. Although $b_3$ and $b_4$ also present a quartic dependence on the Yukawa couplings, such dependence is absent from $b_3 + b_4$. As may be easily proved using Eq.~(\ref{mijl}), and following the same procedure as in section 2, in the case of no mixing the contribution of the higher-order operator to the \lq 22' (\lq 11') Higgs mass matrix elements allows us to replace the factors $m_Q^2$ and $m_U^2$ ($m_Q^2$ and $m_D^2$) in the leading order expressions by $m_{Q}^2 + \overline{m}_{t}^2$ and $m_U^2 + \overline{m}_{t}^2$ ($m_Q^2 + m_b^2$ and $m_D^2 + m_b^2$). The same occurs with the D-term contributions proportional to the square of products of the weak couplings and the Yukawa couplings. It is hence convenient to define \begin{eqnarray} t_Q^U & = & \log\left(\frac{m_Q^2 + \overline{m}_{t}^2}{\overline{m}_{t}^2}\right) \;\;\;\;\; {\rm and} \;\;\;\; t_U = \log\left(\frac{m_U^2 + \overline{m}_{t}^2}{\overline{m}_{t}^2}\right), \nonumber\\ t_Q^D & = & \log\left(\frac{m_Q^2 + m_b^2}{m_b^2}\right) \;\;\;\;\; {\rm and} \;\;\;\; t_D = \log\left(\frac{m_D^2 + m_b^2}{m_b^2}\right) \;\;\;\;\; \label{tqtu} \end{eqnarray} while $t_{QU} = t_Q^U - t_U$ and $t_{QD} = t_Q^D - t_D$. Observe that in the denominators of the expressions for $t_D$ and $t_Q^D$ we have written $m_b$ instead of $\overline{m}_{t}$ since we know that, upon consideration of the whole supersymmetric contributions to the physical masses, radiative corrections should vanish in the supersymmetric limit; we should thus use expansion parameters $t_Q^D$ and $t_D$, which vanish in the limit $m_Q,\; m_D\rightarrow 0$. This change has negligible effects on the Higgs mass computation. In the general case, the way to proceed to obtain the Higgs mass matrix elements at the scale $\overline{m}_{t}$ is the following: the CP-even Higgs mass matrix elements may be decomposed in three terms, namely: \begin{eqnarray} {\cal{M}}^2_{ij}(\overline{m}_{t}) & = & \overline{M}^2_{ij}(\overline{m}_{t}) + \left(\widetilde{{\cal{M}}}^2_{ij} (M_{st}) \right)_{\widetilde{t}} \left(\xi_{i}^{\widetilde{t}}(\overline{m}_{t}) \xi_j^{\widetilde{t}}(\overline{m}_{t})\right)^{-1} \nonumber\\ & + & \left(\widetilde{{\cal{M}}}^2_{ij}(M_{sb}) \right)_{\widetilde{b}} \left(\xi_{i}^{\widetilde{b}}(\overline{m}_{t}) \xi_j^{\widetilde{b}}(\overline{m}_{t})\right)^{-1}, \label{m2ijmt} \end{eqnarray} where $\xi_{i}(\overline{m}_{t})$ denote the anomalous dimension factors, \begin{eqnarray} \xi_1^{\widetilde{b}}(\overline{m}_{t}) = 1 + \frac{3 h_b^2}{32 \pi^2} \; t^{\widetilde{b}}_1 \;\;\;\; && \;\;\;\;\;\;\; \xi_2^{\widetilde{b}} (\overline{m}_{t})= 1 + \frac{3 h_t^2}{32 \pi^2} \; t^{\widetilde{b}}_1 \nonumber\\ \xi_1^{\widetilde{t}}(\overline{m}_{t}) = 1 + \frac{3 h_b^2}{32 \pi^2} \; t^{\widetilde{t}}_1 \;\;\;\; && \;\;\;\;\;\;\; \xi_2^{\widetilde{t}} (\overline{m}_{t})= 1 + \frac{3 h_t^2}{32 \pi^2} \; t^{\widetilde{t}}_1 \label{anomdij} \end{eqnarray} where for convenience we define, \begin{equation} t^{\widetilde{b}}_1 = {\rm Max}(t_Q^D,t_D) \;\;\;\;\;\;\;\;\; t^{\widetilde{t}}_1 = {\rm Max}(t_Q^U,t_U), \end{equation} and also, for later use, \begin{equation} t^{\widetilde{b}}_2 = {\rm Min}(t_Q^D,t_D) \;\;\;\;\;\;\;\;\; t^{\widetilde{t}}_2 = {\rm Min}(t_Q^U,t_U). \label{tbtt} \end{equation} In the above, $\left(\widetilde{{\cal{M}}}^2_{ij}(M_{st})\right)_{\widetilde{t}}$ $\left(\left(\widetilde{{\cal{M}}}^2_{ij}(M_{sb}) \right)_{\widetilde{b}}\right)$ is the contribution to the mass matrix elements coming from the terms frozen at the scale $M_{st}$ ($M_{sb}$), where the stops (sbottoms) are decoupled, with $M_{st}^2 = {\rm Max}(m_Q^2 +\overline{m}_{t}^2,m_U^2+\overline{m}_{t}^2)$ $\left(M_{sb}^2 = {\rm Max}(m_Q^2 + m_b^2,m_D^2+ m_b^2)\right)$, and $\overline{M}^2_{ij}$ are obtained from the mass matrix elements $M^2_{ij}$, Eq.~(\ref{mijl}), by considering the one- and two-loop leading logarithm contributions in the renormalizable Higgs quartic coupling expressions. For example, in the case $m_Q > m_U > m_D$, these contributions to the quartic couplings $\overline{\lambda}_i$ are given by \begin{equation} \overline{\lambda}_i = \lambda_i - \left(\lambda_i(m_Q) - \lambda_i^{{\rm tree}}(m_Q) \right) \left( 1 - a_i(\overline{m}_{t}) \; \widetilde{t}_{Q} \right) , \end{equation} where the $\lambda_i$ are given in Eq.~(\ref{lambimt}), $\lambda_1^{{\rm tree}} = \lambda_2^{{\rm tree}} = -\left(\lambda_3^{{\rm tree}}+\lambda_4^{{\rm tree}}\right) = \left(g^2 + g'^2\right)/4$, and, in order to simplify the presentation, we include all D-term contributions in the definition of the quartic couplings $\overline{\lambda}_i$. Observe that, since $\overline{M}_{ij}^2(\overline{m}_{t})$ contains only the one- and two-loop leading logarithm expressions independent of the mixing mass terms, the quartic couplings $\lambda_5$, $\lambda_6$ and $\lambda_7$ give no contribution to $\overline{M}_{ij}^2(\overline{m}_{t})$. Replacing the quartic coupling $\beta$-functions by their dominant Yukawa coupling dependence, we obtain, in the general case \begin{eqnarray} \overline{\lambda}_1 &=& \frac{g^2 + g'^2}{4} \left[ 1 - \frac{3}{8 \pi^2} h_b^2 \;t^{\widetilde{b}}_1 \right] \nonumber\\ & + & \frac{3}{16 \pi^2}\; h_b^4\; \left(t^{\widetilde{b}}_1 - t^{\widetilde{b}}_2 \right)\; \left[ 1 + \frac{1}{16 \pi^2} \left( \frac{3}{2} \;h_b^2 + \frac{1}{2}\;h_t^2 - 8\; g_3^2 \right) \left(t^{\widetilde{b}}_1 + t^{\widetilde{b}}_2 \right) \right] \nonumber\\ & + & \frac{3}{8 \pi^2}\; h_b^4\; t^{\widetilde{b}}_2 \left[ 1 + \frac{1}{16 \pi^2} \left( \frac{3}{2} \;h_b^2 + \frac{1}{2}\;h_t^2 - 8\; g_3^2 \right) t^{\widetilde{b}}_2 \right] + \Delta_1^D, \label{lambda1} \end{eqnarray} \begin{eqnarray} \overline{\lambda}_2 &=& \frac{g^2 + g'^2}{4} \left[ 1 - \frac{3}{8 \pi^2} h_t^2 \; t^{\widetilde{t}}_1 \right] \nonumber\\ & + & \frac{3}{16 \pi^2}\; h_t^4\; \left(t^{\widetilde{t}}_1 - t^{\widetilde{t}}_2 \right) \; \left[ 1 + \frac{1}{16 \pi^2} \left( \frac{3 \;h_t^2}{2} + \frac{h_b^2}{2} - 8\; g_3^2 \right)\left( t^{\widetilde{t}}_1 + t^{\widetilde{t}}_2 \right) \right] \nonumber\\ & + & \frac{3}{8 \pi^2}\; h_t^4\;t^{\widetilde{t}}_2 \; \left[ 1 + \frac{1}{16 \pi^2} \left( \frac{3 \;h_t^2}{2} + \frac{h_b^2}{2} - 8\; g_3^2 \right) t^{\widetilde{t}}_2 \right] + \Delta_2^D, \label{lambda2} \end{eqnarray} \begin{eqnarray} \overline{\lambda}_3 + \overline{\lambda}_4 & = & \Delta_3^D + \Delta_4^D \nonumber\\ & - & \frac{g^2 + g'^2}{4} \left[ 1 - \frac{3}{16 \pi^2} h_b^2 \; t^{\widetilde{b}}_1 - \frac{3}{16 \pi^2} h_t^2 \; t^{\widetilde{t}}_1 \right], \label{lambda4} \end{eqnarray} where we have already performed the top and bottom mass resummations, leading to the change $\widetilde{t}_X \rightarrow t_X$, with $X = Q,U,D$. In the above, $h_t$ and $h_b$ denote the top and bottom quark Yukawa couplings at the scale $\overline{m}_{t}$ and $g_3$ is the strong gauge coupling at the same scale. The $\Delta_i^D$ terms are the additional leading-log D-term contributions appearing through the $\beta$-functions of the quartic couplings, which contain additional terms proportional to the square of the product of weak gauge couplings and Yukawa couplings. Their two-loop leading-log contributions are very small and can be ignored. Interestingly enough, once these terms are considered together with the terms written explicitly in Eqs. (\ref{lambda1})--(\ref{lambda4}), one recovers the expressions for the D-terms first found in Ref.~\cite{B}. Defining $\Delta_3^D = \left( \Delta_3^D \right)_U + \left( \Delta_3^D \right)_D$, for $m_Q \geq m_D$, we obtain \begin{eqnarray} \Delta_1^D & =& \frac{1}{16 \pi^2} g'^2 h_b^2 \; t_{QD} \nonumber\\ \left(\Delta_3^D\right)_D & =& - \frac{1}{32 \pi^2} g'^2 h_b^2 \; t_{QD} \end{eqnarray} while for $m_D \geq m_Q$ we get \begin{eqnarray} \Delta_1^D & =& -\frac{3}{32 \pi^2} \left(\frac{g'^2}{3} + g^2\right) h_b^2 \; t_{QD} \nonumber\\ \left(\Delta_3^D\right)_D & =& - \frac{3}{64 \pi^2} \left( g^2 - \frac{g'^2}{3} \right) h_b^2 \; t_{QD} \\ \Delta_4^D & = & \frac{3}{32 \pi^2} g^2 h_b^2 \; t_{QD} . \nonumber \end{eqnarray} Analogously, for $m_Q \geq m_U$ we obtain \begin{eqnarray} \Delta_2^D & =& \frac{1}{8 \pi^2} g'^2 h_t^2 \; t_{QU} \nonumber\\ \left(\Delta_3^D \right)_U & =& - \frac{1}{16 \pi^2} g'^2 h_t^2 \; t_{QU} \end{eqnarray} while for $m_U \geq m_Q$ we get \begin{eqnarray} \Delta_2^D & =&- \frac{3}{32 \pi^2} \left(-\frac{g'^2}{3} + g^2\right) h_t^2 \; t_{QU} \nonumber\\ \left(\Delta_3^D\right)_U & =& -\frac{3}{64 \pi^2} \left( g^2 + \frac{g'^2}{3} \right) h_t^2 \; t_{QU} \\ \Delta_4^D & = & \frac{3}{32 \pi^2} g^2 h_t^2 \; t_{QU} . \nonumber \end{eqnarray} Finally notice that $\tan\beta$ is fixed at the scale $m_A$, for $m_A\leq \overline{m}_{t}$, while for $m_A\geq\overline{m}_{t}$, $\tan\beta$ is given by \begin{equation} \label{tanbeta} \tan\beta(\overline{m}_{t})=\tan\beta(m_A)\left[ 1+\frac{3}{32\pi^2}(h_t^2-h_b^2)\log \frac{m_A^2}{\overline{m}_{t}^2} \right] . \end{equation} For the case in which the CP-odd Higgs mass $m_A$ is lower than $M_S = {\rm Max} \left(M_{st},M_{sb}\right)$, we should have decoupled (strictly speaking) the heavy Higgs doublet at the scale $m_A$, and defined an effective quartic coupling for the light Higgs as $\lambda(m_A)=m_h(m_A)/2v^2$, the low energy value of it being obtained by the running of the Standard Model renormalization group equations from the scale $m_A$ to $\overline{m}_t$ \cite{CEQW}. For simplicity we have ignored, in our analytical approximation, the effect of the heavy Higgs doublet decoupling at the intermediate scale. We partially compensate this effect by relating the value of $\tan\beta$ at the scale $\overline{m}_t$ with its corresponding value at the scale $m_A$ through its renormalization-group running, Eq.~(\ref{tanbeta}). The Higgs mass matrix elements at the scales $M_{st},M_{sb}$ may be inferred from the second derivative of the one-loop effective potential, Eq.~(\ref{potgeneral}), at the minimum values for the Higgs fields. Consequently, they can be obtained from the expressions given in Refs.~\cite{ERZ,B}, where all parameters should be assumed to be given at the scale at which the matrix element is evaluated, i.e. $M_{st}$ or $M_{sb}$. As we have shown above, the dominant D-term contributions to the Higgs masses are already taken into account in the one-loop leading logarithmic expressions, much as it happens in the case of $m_Q = m_U$, where the dominant D-term contribution comes from the first two terms in Eq.~(\ref{efflamix}). The D-term contribution of $\Delta_{\rm th} \lambda_{\rm eff}$ becomes relevant only for very large mixing and gives a very small correction to the top quark Yukawa coupling effect. For completeness, we shall include all ${\cal{O}}(g^2h_q^2,g'^2h_q^2)$ D-term contributions, where $q = t,b$. In order to obtain the expressions for $\left(\widetilde{{\cal{M}}}^2_{ij}\right)_{\widetilde{t}} \;, \left(\widetilde{{\cal{M}}}^2_{ij}\right)_{\widetilde{b}}$ one must subtract from the expressions of the matrix elements at the scale $M_{st}$, $M_{sb}$, respectively, the contributions coming from the term dependent on the CP-odd Higgs mass and from the leading order logarithmic expressions. This procedure leaves in the matrix elements all the terms that should be frozen at the scales $M_{st}$, $M_{sb}$, including all terms depending on the squark mixing mass parameters. Using the expressions given in Refs.~\cite{ERZ,B}, and Eqs.~(\ref{lambimt})--(\ref{lambda4}), we obtain, \begin{eqnarray} \label{cpevenst} && \left(\widetilde{{\cal{M}}}_{ij}^2(M_{st}) \right)_{\widetilde{t}} = \frac{3}{8 \pi^2 v^2} \left[ \begin{array}{cc} \widetilde{\Delta}_{11}^{\widetilde{t}} + \left(\widetilde{\Delta}'_{11}\right)^{\widetilde{t}} & \widetilde{\Delta}_{12}^{\widetilde{t}} + \left(\widetilde{\Delta}'_{12}\right)^{\widetilde{t}} \\ \widetilde{\Delta}_{12}^{\widetilde{t}} + \left(\widetilde{\Delta}'_{12}\right)^{\widetilde{t}} & \widetilde{\Delta}_{22}^{\widetilde{t}} + \left(\widetilde{\Delta}'_{22}\right)^{\widetilde{t}} \end{array} \right], \end{eqnarray} \begin{eqnarray} \label{cpevensb} && \left(\widetilde{{\cal{M}}}_{ij}^2 (M_{sb}) \right)_{\widetilde{b}} = \frac{3}{8 \pi^2 v^2} \left[ \begin{array}{cc} \widetilde{\Delta}_{11}^{\widetilde{b}} + \left(\widetilde{\Delta}'_{11}\right)^{\widetilde{b}} & \widetilde{\Delta}_{12}^{\widetilde{b}} + \left(\widetilde{\Delta}'_{12}\right)^{\widetilde{b}} \\ \widetilde{\Delta}_{12}^{\widetilde{b}} + \left(\widetilde{\Delta}'_{12}\right)^{\widetilde{b}} & \widetilde{\Delta}_{22}^{\widetilde{b}} + \left(\widetilde{\Delta}'_{22}\right)^{\widetilde{b}} \end{array} \right], \end{eqnarray} where \begin{eqnarray} \label{delta11b} \widetilde{\Delta}_{11}^{\widetilde{b}}& = & \frac{m_b^4}{\cos^2\beta} \left[\log \left(\frac{m_{\;\widetilde{b}_1}^2 m_{\;\widetilde{b}_2}^2} {\left(m_Q^2 + m_b^2\right)\left(m_D^2 + m_b^2\right)}\right) +\frac{2 A_b(A_b-\mu\tan\beta)} {m_{\;\widetilde{b}_1}^2-m_{\;\widetilde{b}_2}^2} \log\frac{m_{\;\widetilde{b}_1}^2}{m_{\;\widetilde{b}_2}^2}\right] \nonumber \\ & + & \frac{m_b^4}{\cos^2\beta} \left[\frac{A_b(A_b-\mu\tan\beta)} {m_{\;\widetilde{b}_1}^2-m_{\;\widetilde{b}_2}^2} \right]^2 g(m_{\;\widetilde{b}_1}^2,m_{\;\widetilde{b}_2}^2), \end{eqnarray} \begin{eqnarray} \label{delta11t} \widetilde{\Delta}_{11}^{\widetilde{t}}& = & \frac{m_t^4}{\sin^2\beta} \left[\frac{\mu(-A_t+\mu\cot\beta)} {m_{\;\widetilde{t}_1}^2-m_{\;\widetilde{t}_2}^2} \right]^2 g(m_{\;\widetilde{t}_1}^2,m_{\;\widetilde{t}_2}^2), \end{eqnarray} \begin{eqnarray} \label{delta22t} \widetilde{\Delta}_{22}^{\widetilde{t}} & = & \frac{m_t^4}{\sin^2\beta} \left[\log \left(\frac{m_{\;\widetilde{t}_1}^2 m_{\;\widetilde{t}_2}^2} {\left(m_Q^2 + m_t^2\right)\left(m_U^2 + m_t^2\right)} \right) +\frac{2 A_t(A_t-\mu\cot\beta)} {m_{\;\widetilde{t}_1}^2-m_{\;\widetilde{t}_2}^2} \log\frac{m_{\;\widetilde{t}_1}^2} {m_{\;\widetilde{t}_2}^2}\right] \nonumber \\ & + & \frac{m_t^4}{\sin^2\beta} \left[\frac{A_t(A_t-\mu\cot\beta)}{m_{\;\widetilde{t}_1}^2- m_{\;\widetilde{t}_2}^2} \right]^2 g(m_{\;\widetilde{t}_1}^2,m_{\;\widetilde{t}_2}^2), \end{eqnarray} \begin{eqnarray} \label{delta22b} \widetilde{\Delta}_{22}^{\widetilde{b}} & = & \frac{m_b^4}{\cos^2\beta} \left[\frac{\mu(-A_b+\mu\tan\beta)}{m_{\;\widetilde{b}_1}^2- m_{\;\widetilde{b}_2}^2} \right]^2 g(m_{\;\widetilde{b}_1}^2,m_{\;\widetilde{b}_2}^2), \end{eqnarray} \begin{eqnarray} \label{delta12t} \widetilde{\Delta}_{12}^{\widetilde{t}} & = & \frac{m_t^4}{\sin^2\beta} \frac{\mu(-A_t+\mu\cot\beta)} {m_{\;\widetilde{t}_1}^2-m_{\;\widetilde{t}_2}^2} \left[\log\frac{m_{\;\widetilde{t}_1}^2}{m_{\;\widetilde{t}_2}^2} +\frac{A_t(A_t-\mu\cot\beta)} {m_{\;\widetilde{t}_1}^2-m_{\;\widetilde{t}_2}^2} g(m_{\;\widetilde{t}_1}^2,m_{\;\widetilde{t}_2}^2)\right], \nonumber\\ \end{eqnarray} \begin{eqnarray} \label{delta12b} \widetilde{\Delta}_{12}^{\widetilde{b}} & = & \frac{m_b^4}{\cos^2\beta} \frac{\mu(-A_b+\mu\tan\beta)} {m_{\;\widetilde{b}_1}^2-m_{\;\widetilde{b}_2}^2} \left[\log\frac{m_{\;\widetilde{b}_1}^2}{m_{\;\widetilde{b}_2}^2} +\frac{A_b(A_b-\mu\tan\beta)} {m_{\;\widetilde{b}_1}^2-m_{\;\widetilde{b}_2}^2} g(m_{\;\widetilde{b}_1}^2,m_{\;\widetilde{b}_2}^2)\right], \nonumber\\ \end{eqnarray} and \begin{eqnarray} \left(\widetilde{\Delta}'_{11}\right)^{\widetilde{b}} & = & M_Z^2 \left[ 2 m_b^2 f^{\;\widetilde{b}}_1 - m_b A_b f^{\;\widetilde{b}}_2 \right], \;\;\;\;\;\;\;\;\;\; \left(\widetilde{\Delta}'_{11}\right)^{\widetilde{t}} = M_Z^2 m_t \mu \cot\beta f^{\;\widetilde{t}}_2\; \nonumber\\ \left(\widetilde{\Delta}'_{22}\right)^{\widetilde{t}} & = & M_Z^2 \left[ -2 m_t^2 f^{\;\widetilde{t}}_1 + m_t A_t f^{\;\tilde{t}}_2 \right], \;\;\;\;\;\;\;\;\; \left(\widetilde{\Delta}'_{22}\right)^{\widetilde{b}} = - M_Z^2 m_b \mu \tan\beta f^{\;\widetilde{b}}_2\; \nonumber\\ \left(\widetilde{\Delta}'_{12}\right)^{\widetilde{t}} & = & M_Z^2 \left[ m_t^2 \cot\beta f^{\;\widetilde{t}}_1 - m_t \frac{ A_t \cot\beta + \mu}{2} f^{\;\widetilde{t}}_2 \right], \nonumber\\ \left(\widetilde{\Delta}'_{12}\right)^{\widetilde{b}} & = & M_Z^2 \left[ - m_b^2 \tan\beta f^{\;\widetilde{b}}_1 + m_b \frac{A_b \tan\beta + \mu}{2} f^{\;\widetilde{b}}_2 \right], \end{eqnarray} where \begin{equation} g(m_1^2,m_2^2)=2-\frac{m_1^2+m_2^2}{m_1^2-m_2^2} \log\frac{m_1^2}{m_2^2}, \end{equation} \begin{eqnarray} f^{\;\widetilde{t}}_1 & = & \frac{m_Q^2 - m_U^2} {m_{\;\widetilde{t}_1}^2 - m_{\;\widetilde{t}_2}^2} \left(\frac{1}{2} - \frac{4}{3} \sin^2\theta_W\right) \log \left( \frac{m_{\;\widetilde{t}_1}}{m_{\;\widetilde{t}_2}}\right) \nonumber\\ &+& \left(\frac{1}{2} - \frac{2}{3} \sin^2\theta_W \right) \log\left(\frac{m_{\;\widetilde{t}_1} m_{\;\widetilde{t}_2}}{m_Q^2 + m_t^2} \right) \\ & + & \frac{2}{3} \sin^2\theta_W \log \left( \frac{m_{\;\widetilde{t}_1}m_{\;\widetilde{t}_2}} {m_U^2 + m_t^2} \right) \nonumber, \end{eqnarray} \begin{eqnarray} f^{\;\widetilde{b}}_1 & = & \frac{m_Q^2 - m_D^2} {m_{\;\widetilde{b}_1}^2 - m_{\;\widetilde{b}_2}^2} \left(-\frac{1}{2} + \frac{2}{3} \sin^2\theta_W\right) \log \left( \frac{m_{\;\widetilde{b}_1}}{m_{\;\widetilde{b}_2}}\right) \nonumber\\ &+& \left(-\frac{1}{2} + \frac{1}{3} \sin^2\theta_W \right) \log\left(\frac{m_{\;\widetilde{b}_1} m_{\;\widetilde{b}_2}}{m_Q^2 + m_b^2} \right) \\ &-& \frac{1}{3} \sin^2\theta_W \log \left( \frac{m_{\;\widetilde{b}_1}m_{\;\widetilde{b}_2}} {m_D^2 + m_b^2} \right)\nonumber, \end{eqnarray} \begin{eqnarray} f^{\;\widetilde{t}}_2 & = & m_t\; \frac{A_t - \mu \cot\beta}{m_{\;\widetilde{t}_1}^2 - m_{\;\widetilde{t}_2}^2} \left[-\frac{1}{2} \log\left( \frac{m_{\;\widetilde{t}_1}^2} {m_{\;\widetilde{t}_2}^2}\right)\right. \nonumber \\ & + &\left. \left(\frac{4}{3} \sin^2\theta_W - \frac{1}{2} \right) \frac{m_Q^2 - m_U^2} {m_{\;\widetilde{t}_1}^2 - m_{\;\widetilde{t}_2}^2} g(m_{\;\widetilde{t}_1}^2, m_{\;\widetilde{t}_2}^2) \right], \end{eqnarray} \begin{eqnarray} f^{\;\widetilde{b}}_2 & = & m_b\; \frac{A_b - \mu \tan\beta}{m_{\;\widetilde{b}_1}^2 - m_{\;\widetilde{b}_2}^2} \left[ \frac{1}{2} \log\left( \frac{m_{\;\widetilde{b}_1}^2} {m_{\;\widetilde{b}_2}^2}\right) \right. \nonumber \\ & + & \left. \left( \frac{1}{2} -\frac{2}{3} \sin^2\theta_W \right) \frac{m_Q^2 - m_D^2} {m_{\;\widetilde{b}_1}^2 - m_{\;\widetilde{b}_2}^2} g(m_{\;\widetilde{b}_1}^2, m_{\;\widetilde{b}_2}^2) \right]. \end{eqnarray} In the above, all terms should be computed at the scale $M_{st}$ or $M_{sb}$ depending on whether they are associated to stop or sbottom contributions and all ${\cal{O}}(g^4,g^2g'^2,g'^4)$ terms are ignored. It is easy to show that, apart from very small terms of order $M_Z^2/(m_{\;\widetilde{t}_1}^2 + m_{\;\widetilde{t}_2}^2)$, Eqs.~(\ref{cpevenst}), (\ref{cpevensb}) vanish in the case of zero squark mixing. It is also straightforward to show that in the limit of large $\tan\beta$, $m_Q = m_U$ and $m_b = \mu =0$, $\widetilde{\Delta}_{12}^{\widetilde{t}} = 0$ and $3 \left(\widetilde{\Delta}_{22}^{\widetilde{t}} + \left(\widetilde{\Delta}_{22}'\right)^{\widetilde{t}} \right)/16\pi^2v^4$ reproduces the expression of $\Delta_{\rm th} \lambda_{\rm eff}$ given in Eq.~(\ref{umbral}). All parameter values in the expressions for the matrix elements $\widetilde{{\cal{M}}}_{ij}^2$ can be expressed in terms of their values at the scale $\overline{m}_{t}$ by making use of the corresponding $\beta$-and $\gamma$-functions. Having computed the renormalization group improved Higgs mass matrix elements at the scale $\overline{m}_{t}$, the neutral CP-even Higgs mass eigenvalues can be easily derived. They read \begin{eqnarray} m^2_{h(H)} &=& \frac{{\rm Tr} {\cal M}^2 \mp \sqrt{({\rm Tr} {\cal M}^2)^2 - 4 \det {\cal M}^2}}{2} \label{mhH} \end{eqnarray} where \begin{equation} {\rm Tr}{\cal M}^2 = {\cal{M}}_{11}^2 + {\cal{M}}_{22}^2 \;\; ; \;\;\;\;\; \det {\cal M}^2 = {\cal{M}}_{11}^2 {\cal{M}}_{22}^2 - \left( {\cal{M}}_{12}^2 \right)^2 , \label{detm2} \end{equation} and the ${\cal{M}}_{ij}$ are the renormalized Higgs mass matrix elements at the scale $\overline{m}_{t}$, Eq.~(\ref{m2ijmt}). {}From the matrix elements, the mixing angle $\alpha$ is also determined by~\cite{HH}: \begin{eqnarray} \sin 2\alpha = \frac{2{\cal{M}}_{12}^2} {\sqrt{\left({\rm Tr} {\cal M}^2\right)^2-4\det {\cal M}^2}} \end{eqnarray} \begin{eqnarray} \cos 2\alpha = \frac{{\cal{M}}_{11}^2-{\cal{M}}_{22}^2} {\sqrt{\left({\rm Tr} {\cal M}^2\right)^2-4\det {\cal M}^2}} \end{eqnarray} Concerning the running of the squark mass parameters, the dominant contribution comes from gluino-induced effects, which are absent in the case of heavy gluino particles, as we are considering within this work. The remaining contributions are small and have a somewhat complicated dependence on the squark and Higgs spectrum. We shall ignore them within our approximation. We have further defined the light stop and sbottom masses as the values obtained using Eqs. (\ref{masast}), (\ref{masasb}), while taking the running mass parameters at the scale $\overline{m}_{t}$ and adding the QCD-dependent vacuum polarization effects. A more precise definition of the squark masses may be obtained by computing the squark effective potential and adding the full vacuum polarization contributions to the squark masses, much as we have done in the case of the Higgs bosons. We shall concentrate on this subject elsewhere. Figures~3 and 4 show the dependence of the Higgs mass $m_h$ for varying values of $m_Q = A_t = m_A$ and for a fixed value of the right-handed mass parameters $m_U = m_D =1$ TeV and $m_U = m_D = 100$ GeV, respectively. The supersymmetric mass parameter $\mu$ has been set to zero. The solid lines represent the value of the Higgs mass by performing the renormalization group improvement of the effective potential method, as explained in this work. The dotted lines represent the values obtained from the one-loop effective potential, while ignoring the stop decoupling and taking all values to be given at the scale $\overline{m}_{t}$ \cite{ERZ}. The dashed lines are the values obtained for the Higgs mass, while considering that the effective potential is scale-invariant, that is by taking the second derivatives at the scale $M_S$ and rescaling them to the scale $\overline{m}_{t}$ through the appropriate anomalous dimension factors (see e.g. Ref.~\cite{CEQW}). For low values of $m_U$ and $m_D$, as those shown in Fig. 4, the last method becomes accurate for all values of $m_Q < 600$ GeV, while in the second method the departure from the proper renormalization group improved values is faster. Observe that the behaviour shown in Fig. 4 is very similar to the one shown in Fig. 1. Since the main purpose of Figs.~3 and 4 is to compare the results in this paper with other different approaches, based on the one-loop MSSM effective potential, we have plotted in them the running Higgs mass at the scale $\overline{m}_t$, and turned the small D-terms threshold corrections off, i.e. we have put $\widetilde{\Delta}'_{ij}=0$. In the following we shall turn these D-terms on and consider pole Higgs masses. The Higgs masses $m_{h,H,A}$ defined in Eq.~(\ref{mhH}) are all running masses obtained from the effective potential\footnote{In fact, as was noticed in Ref.~\cite{ERZ}, the mass of the CP-odd Higgs, $m_A$, turns out to be scale independent at one loop.}, and evaluated at the top-quark mass scale. To compute the physical (propagator pole) masses $M_{h,H,A}$ one has to correct for the fact that the effective potential is defined at zero external momentum. In fact, the pole and running Higgs masses are related by (see e.g. Ref.~\cite{CEQR}) \begin{equation} \label{polemasas} M^2_{\varphi}=m^2_{\varphi}+{\rm Re}\Delta\Pi_{\varphi}(M^2_{\varphi}) \end{equation} where ${\varphi}=h,H,A$ and \begin{equation} \label{delpolariz} \Delta\Pi_{\varphi}(M^2_{\varphi})=\Pi(M^2_{\varphi})-\Pi(0) , \end{equation} $\Pi_{\varphi}(q^2)$ being the renormalized self-energy of the corresponding Higgs boson. We have computed the Higgs self-energies, at the one-loop level, from the top/stop and bottom/sbottom sectors. The corresponding expressions can be found in Appendix A. Figures~5--7 show the variation of the pole Higgs mass $M_h$ as a function of $m_Q$ for fixed $m_A$ and several fixed values of $m_U$ and of the stop mixing parameter $A_t$. Although in general, for a fixed moderate value of $m_U$ and a fixed value of $A_t$, the Higgs mass increases together with $m_Q$, for large values of $A_t$ and moderate values of $m_U$ or for small values of $m_U$ and moderate values of $A_t$, situations can be observed for which the Higgs mass decreases with larger values of $m_Q$. An understanding of this effect may be obtained by making use of the approximation of Ref.~\cite{CEQW}: although $\log(M_{\rm SUSY}/\overline{m}_{t})$, with $M_{\rm SUSY}^2 \equiv \left(m_{\;\widetilde{t}_1}^2 + m_{\;\widetilde{t}_2}^2 \right)/2$, increases together with the squark-mass parameters, leading to larger values of the Higgs mass, the squark mixing contributions are maximized for values of $A_t^{\rm max} \simeq 2.4 M_{\rm SUSY}$. Hence, for fixed values of $A_t$ a delicate balance between these two effects should be present in order to maximize the Higgs mass. Observe also that for small values of $m_U$, such as the ones analysed in Fig.~7, the configurations which maximize the lightest CP-even Higgs mass $M_h$ correspond to situations for which one of the stops may become light. Indeed, the curves are cut since we have introduced the (crude) experimental constraint $m_{\;\widetilde{t}_{2}}\stackrel{>}{{}_\sim} 45$ GeV on them. An important effect arising in the case of small values of one of the mass parameters, for instance $m_U^2 \simeq M_Z^2$, and $m_Q^2 \gg m_U^2$ as shown in Fig.~7, is that for the same value of $M_{\rm SUSY}$, the maximal value for the Higgs mass is always lower than in the case of $m_Q = m_U$. This is due to the fact that the requirement of having stop masses above the present experimental bound implies $A_t \leq m_Q \simeq \sqrt{2} M_{\rm SUSY}$. Hence, $A_t$ is always significantly lower than $A_t^{\rm max}$, which, as explained above, is the value that maximizes the Higgs mass for that particular value of the average stop mass scale $M_{\rm SUSY}$. One could enquire about the stop and sbottom vacuum polarization contributions in the pole Higgs boson mass definitions. We have checked that these contributions do not give a significant effect in the determination of the neutral Higgs boson masses, unless one of the squarks becomes light (i.e. $\;\widetilde{t}$ and/or $\;\widetilde{b}$) and its couplings to the Higgs fields (i.e. $A_t$, $A_b$, $\mu$) are large. This behaviour is illustrated in Fig.~8 where we plot the neutral Higgs boson running ($m_h$, $m_H$, $m_A$) masses (dotted lines) and pole ($M_h$, $M_H$, $M_A$) masses (solid lines) as functions of $A_t$, for fixed values of the other supersymmetric parameters. We also plot the mass of the lightest stop $m_{\;\widetilde{t}_{2}}$ (dashed line). We see that the departure between the running masses and the pole masses occurs in all cases for large values of $A_t$ and small values of $m_{\;\widetilde{t}_{2}}$. The dependence of the Higgs mass on $m_D$ becomes relevant only for large values of $\tan\beta$. In Fig.~9 we show the variation of $M_h$ as a function of $m_D$ for $\tan\beta=60$, fixed values of $m_Q$, $m_U$ and $m_A$, and two values of $\mu$: $\mu=1$ TeV (solid curves) and $\mu=2$ TeV (dashed curves). The radiative corrections to $M_h$ induced by the bottom/sbottom propagation are always of negative sign and become only relevant for very large values of the $\mu$-parameter. The behaviour with $\mu$ can be understood from the fact that the larger $\mu$, the larger mixing in the sbottom sector and hence lighter sbottom masses are obtained. In this example, we have chosen $m_b(\overline{m}_{t}) = 3$ GeV, which corresponds to a bottom quark pole mass $M_b \simeq 5$ GeV. All the curves are cut by the experimental constraint $m_{\;\widetilde{b}_{2}}\stackrel{>}{{}_\sim} 45$ GeV. Finally Fig.~9 also exhibits the dependence of the Higgs mass on the parameter $A_t$, its dependence with respect to $A_b$ being tiny. \setcounter{equation}{0{Conclusions} We have presented a renormalization group improvement of the effective potential computation of the neutral Higgs masses in the MSSM. The method provides the first calculation of two-loop leading order corrections to the Higgs masses valid for any value of the soft supersymmetry breaking squark mass parameters, $m_Q$, $m_U$, $m_D$, $A_t$ and $A_b$, the CP-odd mass $m_A$, the supersymmetric Higgs mass $\mu$ and $\tan\beta$. This generalization is essential for the computation of the Higgs masses and mixing angles in the presence of light squarks. Our method uses explicit decoupling of stops and sbottoms at their corresponding mass scales, leaving threshold effects in the effective potential (and coupling constants) frozen at the decoupling scales and evolving, in the squared mass matrix, with the anomalous dimensions of the Higgs fields. The threshold effects achieve a complete matching of the effective potential for scales above and below the decoupling scales, and include all higher order (non-renormalizable) terms arising from the whole MSSM effective potential. The effect of considering non-renormalizable threshold effects in the effective potential is twofold: on the one hand it triggers resummations in the renormalization group expansion of the parameters, leading to `physical' expansion parameters; on the other hand, it enables to consider the general case of arbitrary left--right squark mixing, as well as general left- and right-handed soft supersymmetry breaking squark masses. We have corrected the running neutral Higgs boson masses with one-loop self-energy diagrams, where top- and bottom-quarks, stops and sbottoms propagate, to define the corresponding pole masses. The numerical effect of polarizations is relevant only under special circumstances: light squarks and large mixing. We have analysed the general pattern of Higgs masses for general values of the supersymmetric parameters. We have found regions in the parameter space where the radiative corrections become large and negative. They are characterized by large values of the mixing-mass parameters, where the stability of the electroweak minimum can be endangered by the presence of charge and color breaking minima. Our results also allow the evaluation of the relevant radiatively corrected Higgs couplings through the corresponding value of the Higgs angle $\alpha$ \cite{CEQW}. We have neglected, throughout the whole calculation, the possible contribution coming from light charginos/neutralinos. Their effect can be easily included in the threshold terms, as well as in the running of the $\beta$- and $\gamma$-functions, where they appear as ${\cal O}(g^4,g^2g'^2,g'^4)$ terms and are thus numerically unimportant. In Ref.~\cite{CEQW} we have shown that, in the case of a heavy supersymmetric spectrum, our analytical expressions reproduced the Higgs mass spectrum with an error of less than 2--3 GeV. It can be easily checked that light charginos and neutralinos can increase the Higgs mass in $\stackrel{<}{{}_\sim}$ 2--3 GeV \cite{topc,Marc}, with respect to heavy ones, which is indeed within the errors of our different approximations. Nevertheless for completeness, we include in appendix B the leading-log ${\cal{O}}(g^4,g'^2g^2,g'^4)$ chargino and neutralino contributions to the CP-even Higgs masses and the mixing angle. Throughout this work, we have also implicitly assumed that the gluino masses are of order $M_S$. If the gluinos were, instead, much lighter than the characteristic squark masses, the running of the third generation Yukawa couplings would be different, inducing a small indirect effect on the two-loop Higgs mass computation. The third generation Yukawa coupling running is also modified by the presence of light charginos/neutralinos in the spectrum. These two loop contributions to the CP-even Higgs masses are also presented in appendix B. Our present analysis reproduces, with a high level of accuracy, the values of the Higgs masses and mixing angles, for the previously studied case of degenerate left- and right-handed squark mass parameters, and for values of the squark left--right mixing mass parameters lower than the ones giving the maximal values of the lightest CP-even Higgs mass. This comparison holds up to a tiny difference coming from the inclusion in this work of the small D-term threshold contributions and vacuum-polarization effects. This confirms previous results on the upper bound on the lightest Higgs boson mass in the MSSM. \section*{Acknowledgements} We would like to thank A. Brignole, J.R. Espinosa, H. Haber, S. Peris and F. Zwirner for interesting discussions. \newpage
\section*{Acknowledgments} I thank S. Kuhlmann, J. Huston, J. Linnemann, T. Ferbel, and S. Snyder for information about the experiments. This work was supported in part by U.S. National Science Foundation grant number PHY-9507683.
\section{Introduction} \label{sec 1} Let $K$ be any compact, connected set in the plane. The complement of $K$ has one unbounded component and its topological boundary is called the {\em frontier} of $K$, denoted $\mbox {\rm \small frontier}(K)$. The example we are most interested in is when $K$ is the range of a planar Brownian motion run for a finite time (see Figure~\ref{outer}). In this case, Mandelbrot (1982) conjectured that the Hausdorff dimension $\dim (\mbox {\rm \small frontier}(K))$ is $4/3$. Rigorously, the best proven upper bound on the dimension is $3/2 - 1/(4\pi^2) \approx 1.475$ by Burdzy and Lawler (1990). Burdzy (1989) proved that $\mbox {\rm \small frontier}(K)$ has infinite length; our main result improves this to a strict dimension inequality: \begin{figure} \vbox{ \centerline{ \hbox{ \psfig{figure=path4.ps,height=2.5in} $\hphantom{xxx}$ \psfig{figure=outline4.ps,height=2.5in} }}} \caption{ \label{outer} A Brownian path and its frontier} \end{figure} \begin{th} \label{th main} Let $B[0,t]$ denote the range of a planar Brownian motion, run until time $t>0$. There is an $\epsilon > 0$ such that with probability 1, The Hausdorff dimension $ \, \dim (\mbox{\rm \small frontier}(B[0,1]))$ is at least $1 + \epsilon$. Moreover, with probability 1, $$ \inf_{t>0} \, \inf_V \, \dim \Big(\mbox{\rm \small frontier}(B[0,t]) \cap V \Big) \, \geq \, 1 + \epsilon \, , $$ where the inner infimum is over all open sets $V$ that intersect $\, \mbox{\rm \small frontier}(B[0,t])$. \end{th} \noindent{\em Remarks:} The uniformity in $t$ implies that $ \, \dim (\mbox{\rm \small frontier}(B[0,\tau])) \geq 1 + \epsilon$ almost surely for any positive random variable $\tau$ (which may depend on the Brownian motion). We also note that our proof shows that the frontier can be replaced in the statement of the theorem by the boundary of any connected component of the complement $B[0,t]^c$. (One can also infer this from the statement of the theorem by using conformal invariance of Brownian motion). As explained at the end of Section \ref{sec 6}, The result also extends to the frontier of the planar Brownian bridge (which is a closed Jordan curve by Burdzy and Lawler (1990)). Bishop and Jones (1994) proved that if a compact, connected set is ``uniformly wiggly at all scales'', then it has dimension strictly greater than 1. Here we adapt this to a stochastic setting in which the set is likely to be wiggly at each scale, given the behavior at previous scales. The difficulty is in handling statistical dependence. \noindent{\bf Definitions:} Let $G$ be a compact set in the plane with complement $ G^c$, and let $\eta>0$. Denote by \bcorG the set $\{ z \in G : \mbox{\rm dist} (z, G^c) > \eta \cdot \mbox{\rm diam}(G) \}$. Say that the compact set $K \, $ \mbox{\bf $\, \eta$-surrounds} $G$ if $K$ topologically separates \mbox{{\rm core}$(G,\eta)$ } from $ G^c$, i.e., if \mbox{{\rm core}$(G,\eta)$ } is disjoint from the unbounded component of $(K \cap G)^c$. \begin{figure} \centerline{ \psfig{figure=gen4.ps,height=2.5in} } \caption{ \label{gen4} The Gosper island } \end{figure} \begin{th} \label{th any K} Let $G_0$ be the {\bf Gosper Island}, defined in the next section and illustrated in Figure~\ref{gen4}. There exists an absolute constant $\eta_0 > 0$ with the following property. Suppose that $ c_0 >0$, and $K$ is a random compact connected subset of the plane such that for all homothetic images $G=z+r G_0$ of $G_0$ with $r \in (0, 1)$ and $z$ in the plane: \begin{equation} \label{eq gosper-wiggly} {\bf{P}} \Big [ K \, \mbox{ \rm $\eta_0$-surrounds } G \, \mbox { \rm \underbar{or} } \; \, K \cap \crG0 = \emptyset \, \, \Big| \, \, \sigma(K \setminus G^{\circ} ) \Big ] > c_0, \end{equation} where the conditioning is on the $\sigma$-field generated by the random set $K$ outside the interior $G^{\circ}$ of $ G$. Then there is an $\, \epsilon > 0 $, depending only on $c_0$, such that $$ \dim (\mbox {\rm \small frontier}(K)) \geq 1 + \epsilon $$ with probability 1. More generally, $\dim (\mbox {\rm \small frontier}(K) \cap V) \geq 1 + \epsilon$ for any open $V$ intersecting $\mbox {\rm \small frontier}(K)$. \end{th} \begin{figure} \centerline{ \psfig{figure=Surr+Miss.ps,height=2.5in} } \caption{ \label{Surr+Miss} A Brownian motion which surrounds the core and one which misses it. } \end{figure} \noindent{\sc Remarks: } { \bf 1.} In fact, the proof in Section $ 3 \,$ shows that with probability 1, for any connected component $\Omega$ of $K^c$ and any open $V$ intersecting $\partial \Omega$, there is a John domain $\Omega_{\rm J} \subset \Omega$ with closure $\overline{\Omega_{\rm J}}$ contained in $V$, such that $\dim (\partial \Omega \cap \partial \Omega_{\rm J}) \geq 1 + \epsilon$. \newline {\bf 2.} The constant $\eta_0$ will be chosen in the next section to ensure that no ``macroscopic'' line segment can be wholly contained within a $2 \eta_0$-neighborhood of the Gosper Island's boundary $\partial G_0$. The appearance of the Gosper Island might seem strange at this point, but is explained as follows. The hypothesis on $K$ that guarantees ``wiggliness'' should be local to handle dependence (thus it must hold inside each $G$ conditioned on $K \cap G^c$). If $ K \, $ $\, \eta_0$-surrounds $G$, \underbar{or} $\; K \cap \crG0 = \emptyset$, then $\mbox {\rm \small frontier}(K)$ cannot intersect \crG0. Having thus controlled $\mbox {\rm \small frontier}(K)$ inside $G$, away from the boundary of $G$, we must worry about how $\mbox {\rm \small frontier}(K)$ behaves near boundaries of cells $G$, as these run over a partition of the plane. If a small neighborhood of the union of the boundaries of cells $G$ of a fixed size contains no straight line segments of length comparable to $\, \mbox{\rm diam} \, (G) \,$, then no significant flatness can be introduced near cell boundaries. To apply the argument with the same constants on every scale, we need a self-similar tiling where tile boundaries have no straight portions; the Gosper Island yields such a tiling. Proving Theorem \ref{th any K} is the main effort of the paper and is organized as follows. Section~\ref{sec 2} summarizes notation and useful facts about the Gosper Island. We also discuss the notion of a Whitney decomposition with respect to these tiles. Section~\ref{sec 3} constructs a random tree of Whitney tiles for $K$ and reduces Theorem \ref{th any K} to a lower bound on the expected growth rate of the tree, via some general propositions on random trees. In Section~\ref{sec 4} we state a variant of Jones's Traveling Salesman Theorem adapted to the current setting. In Section~\ref{sec 5} this theorem is used to derive the required lower bound on the expected growth rate of the ``Whitney tree'' mentioned above, which then finishes the proof of Theorem~\ref{th any K}. In Section~\ref{sec 6} we verify that the range of planar Brownian motion, killed at an independent exponential time, satisfies the hypothesis of Theorem~\ref{th any K}; this easily yields Theorem \ref{th main}. Finally, section 7 gives a hypothesis on the random set $K$ that is weaker than (\ref{eq gosper-wiggly}), but still implies the conclusion of Theorem \ref{th any K}. \section{Gosper Islands and Whitney tiles} \label{sec 2} The standard hexagonal tiling of the plane is not self-similar, but can be modified to obtain a self-similar tiling. Replacing each hexagon by the union of seven smaller hexagons (of area $1/7$ that of the original -- see Figure \ref{Sub}) yields a new tiling of the plane by $18$-sided polygons; denote by $d_1$ the Hausdorff distance between each of these polygons and the hexagon it approximates. Applying the above operation to each of the seven smaller hexagons yields a $54$-sided polygon with Hausdorff distance $7^{-1/2} \cdot d_1$ from the $18$-sided polygon, which also has translates that tile the plane. Repeating this operation (properly scaled) ad infinitum, we get a sequence of polygonal tilings of the plane, that converge in the Hausdorff metric to a tiling of the plane by translates of a compact connected set $G_0$ called the ``Gosper Island'' (see Gardner (1976) and Mandelbrot (1982)). \begin{figure} \centerline{ \psfig{figure=Sub.ps,height=1.5in} } \caption{ \label{Sub} Substitution defining Gosper island } \end{figure} \begin{figure} \vbox{ \centerline{ \hbox{ \psfig{figure=gen0.ps,height=1.0in} $\hphantom{xxx}$ \psfig{figure=gen1.ps,height=1.0in} $\hphantom{xxx}$ \psfig{figure=gen2.ps,height=1.0in} $\hphantom{xxx}$ \psfig{figure=gen3.ps,height=1.0in} }}} \caption{ \label{Gosper} First four generation of the construction} \end{figure} \noindent{\bf Notation:} We normalize $G_0$ to be centered at the origin and have diameter 1. Denote by ${\cal{D}}_0$ the set of translates of $G_0$ that form a tiling of the plane (depicted in Figure \ref{TileUnion}). This tiling is {\bf self-similar}, i.e., there is a complex number $\lambda$ with $|\lambda| >1$ such that for each tile $G \in {\cal{D}}_0$, the homothetic image $\lambda \cdot G$ is the union of tiles in ${\cal{D}}_0$. (For the tiling by Gosper Islands, $|\lambda| = 7^{1/2}$.) For each integer $n$, we denote by ${\cal{D}}_n$ the scaled tiling $\{ \lambda^{-n} \cdot G \; : \: G \in D_0 \}$, and let $\, {\cal{D}} = \cup_{n=0}^{\infty} {\cal{D}}_n$. If $G \in {\cal{D}}_n$ we say that $G$ is a tile of index $n$ and write $||G||=n$. Every tile $ G \in {\cal{D}}_n$ is contained in a unique tile of ${\cal{D}}_{n-1}$, denoted $\, \mbox{\rm parent}(G)$. Each tile $G$ is centrally symmetric about a ``center point'' $z$; for any $\theta >0$, denote by $\theta \odot G = z+ \theta \cdot (G-z)$ the expansion of $G$ by a factor $\theta$ around $z$. \begin{figure} \centerline{ \psfig{figure=Tiling.ps,height=2.5in} } \caption{ \label{TileUnion} A self-similar tiling of the plane } \end{figure} We record several simple properties of the tiling by Gosper Islands, which will be useful later. \begin{enumerate} \item \label{min-dist} There is some minimal distance $d_0$ between any two nonadjacent tiles of ${\cal{D}}_0$. \item \label{eta} There is an $\eta_0 > 0$ such that any line segment of length $d_0$ must intersect $\mbox{\rm core}(G, 2\eta_0)$ for some $G \in {\cal{D}}_0$. (The existence of $\eta_0$ follows by a compactness argument from the fact that $\partial G_0$ contains no straight line segments.) \item \label{inrad} The Gosper Island $G_0$ contains an open disk centered at the origin which in turn contains $\lambda^{-1} G_0$. \item \label{odot} The blow-up $\lambda^3 \odot D$ contains $\lambda \odot \mbox{\rm parent} (D)$ for any $D \in {\cal{D}}$ (see Figure \ref{WhitTile}). \item \label{neighbor} If $||G|| = ||G'|| - 1$ for neighboring tiles $G$ and $G'$, then $\lambda \odot G$ contains $\lambda \odot G'$. (See Figure~\ref{WhitTile}.) \item \label{union} The blow-up $\lambda \odot G$ is contained in $\bigcup (\lambda \odot G'')$ where the union is over all neighbors $G''$ of $G$ of index $||G||$. \item \label{Jordan} The boundary of $G_0$ is a Jordan curve. To see this note that when we replace each segment or length $r$ by the three segments of the next generation, they remain within distance $r \sqrt{3}/14$ of the segment. Thus the limiting arc is within $$ r \frac {\sqrt {3}}{14} \sum_{n=0}^\infty (\frac 1{\sqrt{7}})^n = {r \sqrt{21}\over 14(\sqrt{7} -1)} \approx r(0.198892),$$ of the segment. If $I_1, I_2, I_3$ are consecutive segments of length $r$ then $\mbox{\rm dist} (I_1, I_3) = r$, so this shows the limiting arcs corresponding to them are at least distance $r/2$ apart. Thus the boundary of the Gosper Island is a Jordan curve, indeed, is the image of the unit circle under a map $f$ satisfying $$ \frac 1C \leq {|f(x) - f(y)| \over |x-y|^\alpha } \leq C,$$ where $\alpha = \frac 12 \log 7/\log 3$. \item \label{annulus} For any $\eta>0$, there is a topological annulus with a rectifiable boundary, which separates $\mbox{\rm core}(G_0,\eta)$ from the boundary $\partial G_0$ of the Gosper island. \newline (By the previous property, the interior $G_0^\circ$ of $G_0$ is simply connected, so this annulus can be obtained, for instance, by applying the Riemann mapping theorem.) \end{enumerate} \noindent{\bf Definitions: } Let $K$ be a compact connected subset of the plane. We say that $\, G \in {\cal{D}} \, $ is a {\bf Whitney~tile} for $\, K \,$ if $\, \lambda \odot G \,$ is disjoint from $\, K$, but $\, \lambda \odot \mbox{\rm parent}(G) \, $ intersects $\, K$. (See Figure~\ref{WhitTile}.) Let $W_K$ denote the set of Whitney tiles for $K$. This collection is called a Whitney decomposition of $K^c$, since it decomposes $K^c$ into a countable union of tiles (disjoint except for their boundaries) each with diameter comparable to its distance from $K$. See Figure \ref{TileChain}. A chain of adjacent tiles $\{ G_1 , G_2 , \ldots , G_j \} \, $ in $\, W_K \,$ such that $\, G_i \subset \lambda^{5} \odot G_1 \, $ and $\, ||G_{i}|| \geq ||G_{i-1}|| \,$ for all $\, i \in \{2, \ldots , j \} \, $ is called a {\bf Whitney~chain} (see Figure~\ref{TileChain}). Given $G \in W_K$, define $W_K^G \subset W_K$ to be the set of tiles $G'$ such that there is a Whitney chain $\{ G_1 , G_2 , \ldots , G_j \}$ with $G_1 = G $ and $G_j = G'$. \begin{figure} \centerline{ \psfig{figure=WhitDefnSim.ps,height=2.5in} } \caption{ \label{WhitTile} Boundary misses $\lambda \odot Q$, but hits $\lambda \odot \mbox{\rm parent}(Q)$. } \end{figure} \begin{figure} \centerline{ \psfig{figure=Dec+Chain.ps,height=2.5in} } \caption{ \label{TileChain} Whitney decomposition and a chain of tiles } \end{figure} Note the following property of the Whitney decomposition, which holds for any connected component $\Omega$ of $K^c$: \begin{equation} \label{Whit-property} \mbox{ \rm For any open $V$ intersecting $\, \partial \Omega\, $, there is a tile } G_* \in W_K \mbox{ with } \lambda^{5} \odot G_* \subset V. \end{equation} \begin{lem} \label{lem 1} If $G_1 , G_2 \in W_K$ are adjacent then $||G_1|| - ||G_2||$ is 0 or $\pm 1$. \end{lem} \noindent{\sc Proof:} Suppose $\, ||G_1|| - ||G_2|| \geq 2$. Let $G$ be the tile of index $\, ||G_2||+1 \,$ that contains $\, G_1$, and observe that $G$ is adjacent to $G_2$. Then by Property \ref{neighbor} of the Gosper tiling, \newline $\lambda \odot \mbox{\rm parent} (G_1) \subset \lambda \odot G \subset \lambda \odot G_2$ and maximality of $G_2$ is violated. $\fbox{\hphantom{} \vphantom{x}}$ \begin{lem} \label{lem 2} Suppose ${\cal{C}} \subset W_K \cap {\cal{D}}_n$ is a collection of tiles whose union topologically surrounds a smaller Whitney tile $G \in W_K \cap {\cal{D}}_{n+k}$ where $k > 0$. Then ${\cal{C}}$ surrounds a a point of $K$. \end{lem} \noindent{\sc Proof:} The $k$-fold parent of $G$ is a tile $D \in{\cal{D}}_n$ which is surrounded by ${\cal{C}}$. Applying Property \ref{union} inductively shows that the union of $\lambda \odot G'$ for $G' \in {\cal{C}}$ surrounds whatever part of $\lambda \odot D$ it does not contain. Thus maximality of $G$ implies that $\lambda \odot D$ intersects $K$, and any point of intersection is surrounded by ${\cal{C}}$. $\fbox{\hphantom{} \vphantom{x}}$ \begin{lem} \label{lem 3} Suppose that $ \, G \in W_K \, $ and that there is a Whitney chain from some larger tile outside $\, \lambda^{5} \odot G \, $ to $\, G$. For any $\, n > ||G|| \, $ define $\, \mbox{\rm Wall} (G,n) \,$ to be the set $ \, W_K^G \cap {\cal{D}}_n$. (See Figure~\ref{TileWall}.) Let $$ E_n = \bigcup \{D \, : \, D \in \mbox{\rm Wall} (G,n) \} \cup \partial (\lambda^{5} \odot G) \, . $$ Then $ \, E_n \,$ is a connected set which topologically separates $\, G \,$ from $\, K \,$. Furthermore, If $\, \Gamma \, $ is a Jordan curve separating $\, G \, $ from the complement of $\, \lambda^{5} \odot G \, $, then every component of $\, \bigcup \{D \, : \, D \in \mbox{\rm Wall} (G,n) \} \,$ intersecting the domain bounded by $\, \Gamma \,$ also intersects $\Gamma$. \end{lem} \begin{figure} \centerline{ \psfig{figure=Wall+Ugh.ps,height=2.5in} } \caption{ \label{TileWall} The sets Wall(G,n) and U(G,h). } \end{figure} \noindent{\sc Proof:} By Lemma \ref{lem 1} any path which connects $G$ to $K$ must hit Whitney tiles of every index larger than $||G||$. Thus any such path either hits $\mbox{\rm Wall}(G,n)$ or must leave $\lambda^{5} \odot G$, proving that $E_n$ separates $G$ from $K$. Connectedness follows from the last assertion of the lemma for $\Gamma = \partial (\lambda^{5} \odot G)$, so it remains only to prove the last assertion. Suppose to the contrary that there is a component $U$ of $\, \bigcup \{D \, : \, D \in \mbox{\rm Wall} (G,n) \} \,$ which intersects the domain bounded by $\Gamma$ but is disjoint from $\Gamma$ itself. The union of all Whitney tiles which are in the unbounded component of $U^c$ and are adjacent to $U$ is a connected set. By Lemma \ref{lem 1} all of these tiles have index $n-1$ or $n+1$. By connectedness and Lemma \ref{lem 1}, they must all have a single index. Suppose they all have index $n+1$. Since tiles in $U$ can be connected to $G$ by Whitney chains which don't cross any tile of index $n+1$, this means $G$ is in a bounded component of the complement of $U$, which contradicts our assumption that $G$ could be connected by a Whitney chain to a larger tile outside $\lambda^{5} \odot G$. Thus the adjacent tiles must all have index $n-1$. But then by Lemma \ref{lem 2} these adjacent tiles must also surround a point of $K$, which implies that $K$ is not connected, another contradiction. $\fbox{\hphantom{} \vphantom{x}}$ The next two lemmas are needed in order to show that if ``major portions'' of a wall of Whitney tiles can be covered by a thin strip, then $K$ must intersect the core of an appropriate tile $G''$ without $\eta$-surrounding it; the latter event is controlled by the hypothesis of Theorem~\ref{th any K}. \begin{lem} \label{lem short segment} Fix $G' \in {\cal{D}}$ and $\beta \in (0, \eta_0)$. Let $\hat U$ be any connected set intersecting both \newline $\partial (\lambda^{5} \odot G')$ and $\lambda^3 \odot G'$. Suppose that $\hat U \cap (\lambda^{5} \odot G')$ is contained in an infinite open strip of width $2 \beta \mbox{\rm diam} (G')$. Then there is a tile $G''$ contained in $\lambda^{5} \odot G'$ and of the same index as $G'$, such that $\hat U$ intersects $\mbox{\rm core}(G'', 2\eta_0-2\beta)$. \end{lem} \noindent{\sc Proof:} Pick a point $x \in \hat U \cap (\lambda^3 \odot G')$ and choose $y \in \hat U \cap \partial (\lambda^4 \odot G)$ connected to $x$ inside $\hat U \cap (\lambda^4 \odot G)$. By Property \ref{min-dist} of the tiling, the segment $\overline{xy}$ has length at least $\, d_0 \cdot \mbox{\rm diam} (G')$, so by Property \ref{eta} of the tiling, there is a tile $G''$ of the same index as $G'$, such that $\overline{xy}$ contains some point $z \in \, \mbox{\rm core}(G'', 2 \eta_0)$. By Property \ref{inrad} and convexity of disks, $z \in \lambda^5 \odot G$, and therefore $G'' \subset \lambda^5 \odot G$. Observe that $\mbox{\rm dist} (z, \hat U) < {2 \beta \mbox{\rm diam} (G')}$, for if not, removing the open disk centered at $z$ of radius $2 \beta \mbox{\rm diam} (G')$ from the infinite strip would contradict the connectedness of $\hat U$. This observation implies the assertion of the lemma. $\fbox{\hphantom{} \vphantom{x}}$ Let $G$ be any Whitney tile with $\lambda^{5} \odot G$ not containing all of $K$. Let $\gamma$ be a circle centered at $\mbox{\rm center} (G)$ which separates $\lambda^4 \odot G$ from $\partial (\lambda^5 \odot G)$. (Such a circle exists by Property \ref{inrad} of the tiling.) For any positive integer $h$, let $U(G,h)$ be the union of all tiles $D \in W_K^G$ of index $||G|| + h$ such that $D$ intersects the disk bounded by $\gamma$ (see Figure~\ref{TileWall}). \begin{lem} \label{lem wiggly -> beta} Choose $a_2$ so that $\lambda^{3 - a_2} \leq \eta_0 / 2$. With $G$ as above, let $G'$ be a tile in $W_K^G$ with $||G|| < ||G'|| < ||G|| + h - a_2$ such that $\lambda^{5} \odot G'$ is contained in the disk bounded by $\gamma$. Suppose that $U (G,h) \cap (\lambda^{5} \odot G')$ is covered by an open strip of width $2 \beta \mbox{\rm diam} G'$ with $\beta < \eta_0 /4$. Then there is a tile $G'' \subset \lambda^{5} \odot G'$ of the same index as $G'$, such that $K$ intersects $\mbox{\rm core} (G'', \eta_0) $ without $\eta_0$-surrounding $G''$. \end{lem} \noindent{\sc Proof:} By Lemma~\ref{lem 3}, $U (G,h) \cup \gamma$ is connected and therefore satisfies the hypotheses of Lemma~\ref{lem short segment}; let $G''$ be a tile as in the conclusion of that lemma. Since $2\eta_0-2\beta > 3 \eta_0/2$, we can pick a point $u$ in $U(G,h) \cap \mbox{\rm core}(G'', 3 {\eta_0}/2)$. This clearly prevents $K$ from $\eta_0$-surrounding $G''$. For any Whitney tile $D$ of index $||G||+h$, the blow-up $\lambda^3 \odot D$ intersects $K$ (by Property \ref{odot} of the tiling). Since $U (G,h) \cap (\lambda^{5} \odot G')$ is a union of tiles of index $||G||+h$, it follows that $$ \mbox{\rm dist}(u,K) < |\lambda|^{3-||G||-h} \leq |\lambda|^{3-a_2-||G'||} \leq {\eta_0 \over 2} \mbox{\rm diam} (G'') \, , $$ by the choice of $a_2$. Therefore $K$ intersects $\mbox{\rm core}(G'',\eta_0)$. $\fbox{\hphantom{} \vphantom{x}}$ \section{A tree of Whitney tiles} \label{sec 3} Fix a compact, connected $K \subset \mathop{\raise .45ex\hbox{${\bf\scriptstyle{|}}$$, a tile $G_* \in W_K$, and a positive integer, $h$. We construct a tree $T = T(K , G_* , h)$ of Whitney tiles. The root of $T$ is $G_*$ and the remaining generations of $T$ are defined recursively as follows. Assume $T$ has been defined up to generation $n$ and for each $G$ in $T_n$, the $n^{th}$ generation of $T$, define $\widetilde T_{n+1} (G)$ to be the set of tiles $D$ with the following properties: \begin{quote} 1. $||D|| = ||G_*|| + (n+1)h$; 2. $D \in W_K^G$; 3. $\lambda^{5} \odot D \subset \lambda^{5} \odot G$. \end{quote} Let $ T_{n+1} (G)$ be a subcollection of $\widetilde T_{n+1} (G)$ which has maximal cardinality among all subcollections ${\cal{C}}$ for which the expanded tiles $\{ \lambda^{6} \odot D : D \in {\cal{C}} \}$ are disjoint. By maximality, $\bigcup\{ \lambda^{7} \odot D \, : \, D \in T_{n+1} (G) \}$ contains all tiles in $\widetilde T_{n+1} (G)$ and therefore \begin{equation} \label{eq:prune} |\widetilde T_{n+1} (G)| \leq |\lambda|^{14} |T_{n+1} (G)| \, . \end{equation} The children of $G$ in $T$ are defined to be the collection $ T_{n+1} (G)$. Some trivial inductive observations are that $T_n \subset {\cal{D}}_{nh + ||G_*||}$, that each $G \in {\cal{D}}_n$ is connected to $G_*$ by a Whitney chain, and that the sets $\lambda^{5} \odot G$ are disjoint as $G$ runs over any $T_n$. \noindent{\bf Some tree terminology:} $\:$ Let $V$ be a countable set. \begin{description} \item{(i)} A mapping {\bf T} from a probability space ${\cal S}$ to the set of trees on the vertex set $V$ is {\bf measurable} with respect to a $\sigma\mbox{-field} $ $ {\cal{F}}$ on ${\cal S}$, if for any pair $\{v,v'\} \subset V$, the event $\left[\{v,v'\} \mbox{\rm is an edge of {\bf T} } \right]$ is in $ {\cal{F}}$. \item{(ii)} For any tree $T$ with vertex set contained in $V$, and any element $v \in V$, define $\mbox{\rm trunc}_v (T)$ to be null if $v$ is not a vertex of $T$, and otherwise let $\mbox{\rm trunc}_v (T)$ be $T$ with the part below $v$ removed; more precisely, the vertices of $\mbox{\rm trunc}_v (T)$ are the vertices of $T$ not separated from the root by $v$, and the edges are the edges of $T$ spanning pairs of vertices in this smaller vertex set. \end{description} For any $G \in {\cal{D}}$, let ${\cal{F}}_G$ denote the $\sigma\mbox{-field}$ generated by the events $\{ D \cap K \neq \emptyset \}$ for all tiles $D$ for which either $||D|| \leq G$ or the interior of $D$ is disjoint from $\lambda^{5} \odot G$. \begin{lem} \label{lem tree-dep} On the event $||G|| = nh + ||G_*||$, the random variable $\mbox{\rm trunc}_G \circ T$ is measurable with respect to ${\cal{F}}_G$. \end{lem} \noindent{\sc Proof:} Suppose $||G|| = nh + ||G_*||$ and consider an event of the form $$\{ \{ D , D' \} \mbox{ is in the edge set of } \mbox{\rm trunc}_G (T) \} ,$$ where $D$ and $D'$ are tiles of index $mh + ||G_*||$ and $(m+1) h + ||G_*||$ respectively, \newline with $\lambda^{5} \odot D' \subset \lambda^{5} \odot D$. If $m \geq n$ and $\lambda^{5} \odot D$ is not disjoint from $\lambda^{5} \odot G$, then the edge $\{ D , D' \}$ cannot be in $\mbox{\rm trunc}_G (T)$. If $m < n$ or $\lambda^{5} \odot D$ is disjoint from $\lambda^{5} \odot G$ then the event that $\{ D , D' \}$ is an edge of $T$ is the union of events witnessed by particular Whitney chains of tiles, all tiles being either disjoint from $\lambda^{5} \odot G$ or of index at most $||G||$, so the event is measurable with respect to ${\cal{F}}_G$. $\fbox{\hphantom{} \vphantom{x}}$ The next lemma requires the traveling salesman theorem described in the next section, so its proof is delayed until Section~\ref{sec 5}. \begin{lem} \label{lem main} Assume the random set $K$ satisfies the hypotheses of Theorem~\ref{th any K}. Fix any tile $G_*$ and $h > 0$ and let $T$ be the random tree $T(K , G_* , h)$. There are constants $c_1 , c_2 > 0$ such that for any tile $G \in {\cal{D}}_{nh+||G_*||}$, \begin{equation} \label{eq main} {\bf{E}} [\# T_{n+1} (G) \| {\cal{F}}_G ] \geq c_1 h |\lambda|^h - c_2 |\lambda|^h \end{equation} on the event that $G_* \in W_K$, the tile $G$ is in $T_n$, and $\lambda^{5} \odot G_*$ does not contain $K$. The constants $c_1$ and $c_2$ depend only on $c_0$. \end{lem} To prove Theorem \ref{th any K}, we also need two general lemmas concerning trees. Define the {\bf boundary} $\partial T$ of the infinite rooted tree $T$ to be the set of infinite self-avoiding paths from the root. The next lemma is implicit in Hawkes (1981) and can be found in a stronger form in Lyons (1990). For convenience, we include the short proof. \begin{lem} \label{lem lyons} Let $T$ be an infinite rooted tree. Given constants $C>0$ and $ \theta>1$, put a metric on $\partial T$ by \begin{equation} \label{eq: metric} \mbox{\rm dist} (\xi , \xi') = C \theta^{-n} \mbox{ if } \xi \mbox{ \rm and } \xi' \mbox{ share exactly } n \mbox{ edges.} \end{equation} Suppose that independent percolation with parameter $p \in (0,1)$ is performed on $T$, i.e., each edge of $T$ is erased with probability $1-p$ and retained with probability $p$, independently of all other edges. If $$ \dim(\partial T) < \alpha = {\log (1/p) \over \log \theta } $$ then with probability 1, all the connected components of retained edges in $T$ are finite. \end{lem} \noindent{\sc Proof:} It suffices to show that the connected component of the root is finite almost surely. For any vertex $v$ of $T$, denote by $|v|$ the number of edges between $v$ and the root. By the dimension hypothesis and the definition of the metric on $\partial T$, there must exist cut-sets $\Pi$ in $T$ for which the $\alpha$-dimensional cut-set sum $$ \sum_{v \in \Pi} \theta^{-|v| \alpha} = \sum_{v \in \Pi} p^{|v|} $$ is arbitrarily small. But for any cutset $\Pi$, the right-hand side is the expected number of vertices in $\Pi$ which are connected to the root after percolation; this expectation bounds the probability that the connected component of the root is infinite. $\fbox{\hphantom{} \vphantom{x}}$ The next lemma formalizes the notion of a random tree which ``stochastically dominates'' the family tree of a branching process. We require the analogue of a filtration in our setting. \noindent{\bf Definition:} $\:$ Let $V$ be a countable set and let $T$ be a random tree with vertex set contained in $V$, i.e., $T$ is a measurable mapping from some probability space $\, \langle {\cal S}, {\cal A}, {\bf{P}} \rangle$ to the set of trees on the vertex set $V$. Say that $\sigma\mbox{-field}$s $\{{\cal{F}}_v \, : \, v \in V\}$ on ${\cal S}$ form a {\bf tree-filtration} if for any $v,w \in V$ and any $A \in {\cal{F}}_v$, the event $A \cap \{ w \mbox{ \rm is a descendant of } v \mbox{ in } T \} \, $ is ${\cal{F}}_w$-measurable. \begin{lem} \label{lem tree} Let $V$ be a countable set and let $T$ be a random tree with vertex set contained in $V$. We assume that $T$ is rooted at a fixed $v_* \in V$. Assume that $b > 1$ and a tree-filtration $\{ {\cal{F}}_v \, : \, v \in V \}$ exists such that $\mbox{\rm trunc}_v (T)$ (defined before Lemma~\ref{lem tree-dep}) is ${\cal{F}}_v$-measurable for each $v \in V $, and the conditional expectation $$ {\bf{E}} ( \mbox{number of children of } v \mbox{ in } T \| {\cal{F}}_v) \geq b . $$ If every vertex of $T$ has at most $M$ children and at least $m$ children, $m \geq 0$, then \begin{enumerate} \item The probability that $T$ is infinite is at least $1-q > 0$, where $q$ is the unique fixed point in $[0,1)$ of the polynomial $$ \psi (s) = s^m + {b-m \over M-m} (s^M - s^m) \, . $$ (Observe that $q = 0$ when $m > 0$.) \item ${\bf{P}} (T \mbox{ is infinite } \| \tilde {\cal{F}}_{v_*} ) \geq 1 - q$ for any $\tilde {\cal{F}}_{v_*} \subset {\cal{F}}_{v_*}$. \item If $\partial T$ is endowed with the metric (\ref {eq: metric}), then $\dim (\partial T) \geq \log b / \log \theta$ with probability at least $1 - q$. \end{enumerate} \end{lem} \noindent{\sc Proof:} \noindent{\bf 1.} Let $ |T_n|$ be the size of the $n^{th}$ generation $T_n$ of $T$. and let $\psi_n (s)$ denote the $n$-fold iterate of $\psi$. We claim that for $s \in [0,1]$, \begin{equation} \label{eq claim 1} {\bf{E}} s^{ |T_n|} \leq \psi_n (s) . \end{equation} When $n = 1$, convexity of $x \mapsto s^x$ implies that $$ s^{|T_1|} \leq s^m + {|T_1| - m \over M-m} (s^M - s^m) $$ and the claim follows by taking expectations: $${\bf{E}} s^{|T_1|} \leq s^m + {{\bf{E}} |T_1| - m \over M-m} (s^M - s^m) \leq \psi_1 (s)$$ since $s^M - s^m \leq 0$. For $n > 1$ proceed by induction. Let $|T_{n+1} (v)|$ be the number of children of $v$ if $v \in T_n$ and zero otherwise, and use the argument from the $n=1$ case to see that ${\bf{E}} (s^{ |T_{n+1} (v)|} \| {\cal{F}}_v ) \leq \psi (s)$ on the event $v \in T_n$ (which is an event in ${\cal{F}}_v$). Giving $V$ an arbitrary linear order (denoted ``$<$''), we have in particular $${\bf{E}} (s^{ |T_{n+1}(v)|} \| T_n \mbox{\rm and } T_{n+1} (w) \mbox { for } w < v \mbox { in } T_n) \leq \psi (s) \, . $$ for $v \in T_n$. Since $$ s^{ |T_{n+1}|} =\prod_{v \in T_n} s^{ |T_{n+1}(v)|} \, , \; \mbox{ this yields } \; \; {\bf{E}} (s^{ |T_{n+1}|} \| T_n) \leq \psi (s)^{ |T_n|} \, . $$ Taking expectations and applying the induction hypothesis with $\psi (s)$ in place of $s$ gives $$ {\bf{E}} s^{ |T_{n+1}|} \leq {\bf{E}} \Big(\psi (s)^{ |T_n|}\Big) \leq \psi_n (\psi (s)) = \psi_{n+1} (s) \, , $$ proving the claim (\ref{eq claim 1}). From (\ref{eq claim 1}) we see that ${\bf{P}} ( |T_n| = 0) \leq E q^{ |T_n|} \leq \psi(q) =q $, establishing the first conclusion of the lemma. \noindent{\bf 2.} By copying the derivation of~(\ref{eq claim 1}), inserting an extra conditioning on $\tilde {\cal{F}}_{v_*}$, one easily verifies that $ {\bf{E}} (s^{ |T_n|} \| \tilde {\cal{F}}_{v_*}) \leq \psi_n (s)$, and the rest of the argument is the same as in the first part. \noindent{\bf 3.} Let $T' (v)$ be the connected component of the subtree of $T$ below $v$ after removing each vertex of $T$ below $v$ independently with probability $1-p$. For $p > 1/b$, let $q_p \in (0,1)$ solve $q_p = 1 + (bp/M) (q_p^M - 1)$. We apply the second part of the lemma to $T' (v)$ conditioned on ${\cal{F}}_v$ to see that $${\bf{P}} (T' (v) \mbox{ is infinite } \| {\cal{F}}_v ) \geq 1 - q_p$$ for $v \in T$. By Lemma \ref{lem lyons}, the event $\{ \dim (\partial T) < |\log p| / \log \theta \}$ is contained up to null sets in the event $\{ T' (v) \mbox{ is finite for all } v \in T_n \}$. Thus \begin{eqnarray*} {\bf{P}} \Big( \dim (\partial T) < |\log p| / \log \theta \, \Big| \, T_n \Big) & \leq & {\bf{P}} \Big( \cap_{v \in T_n} T' (v) \mbox{ finite } \, \Big| \, T_n \Big) \\[2ex] & = & \prod_{v \in T_n} {\bf{P}} \Big( T' (v) \mbox{ finite } \, \Big| \, T_n \, , \, T' (w) \mbox{ finite for all } w < v \mbox{ in } T_n \Big) \\[2ex] & \leq & q_p^{ |T_n|} \end{eqnarray*} since each event conditioned on is in the corresponding ${\cal{F}}_v$. Taking expectations yields $${\bf{P}} \Big( \dim (\partial T) < {|\log p| over \log \theta} \Big) \leq \psi_n (q_p) .$$ Since $q_p < 1$ for each $p > 1/b$, this goes to $q$ as $n \rightarrow \infty$, proving the last conclusion of the lemma. $\fbox{\hphantom{} \vphantom{x}}$ \noindent{\sc Proof of Theorem}~\ref{th any K}: Put a metric on $\partial T$ by $$ \mbox{\rm dist} (\xi , \xi') = |\lambda|^{-(n+1)h - ||G_*||} \mbox{ if } \xi \mbox{\rm and } \xi' \mbox{ share exactly } n \mbox{ edges.} $$ Each $\xi = (G_1 , G_2 , \ldots) \in \partial T$ defines a unique limiting point $\phi (\xi) \in \mbox {\rm \small frontier}(K)$ which is the decreasing limit of the set $\lambda^{5} \odot G_n$. If $\xi = (G_1 , G_2 , \ldots)$ and $\xi' = (G_1' , G_2' , \ldots)$ share exactly $n$ edges, then by definition of $ T_{n+1}$, the expanded tiles $\lambda^{6} \odot G_{n+1}$ and $\lambda^{6} \odot G_{n+1}'$ are disjoint. Since $\phi (\xi) \in \lambda^{5} \odot G_{n+1}$ and $\phi (\xi') \in \lambda^{5} \odot G_{n+1}'$, it follows from Property \ref{min-dist} of the tiling that $$|\phi (\xi) - \phi (\xi')| \geq d_0 |\lambda|^{-(n+1)h - ||G_*||} .$$ Thus $$|\phi (\xi) - \phi (\xi')| \geq d_0 \cdot \mbox{\rm dist}(\xi , \xi')$$ and since the range of $\phi$ is included in $\mbox {\rm \small frontier}(K) \cap \lambda^{5} \odot G_*$ it follows that \begin{equation} \label{eq local dim} \dim (\mbox {\rm \small frontier}(K) \cap \lambda^{5} \odot G_*) \geq \dim (\partial T) . \end{equation} From Lemma \ref{lem tree} and the conclusion of Lemma~\ref{lem main}, we see that \begin{equation} \label{eq dim bound} \dim (\partial T (K , G_* , h)) \geq { \log ( c_1 h |\lambda|^h - c_2 |\lambda|^h ) \over h \log |\lambda|} \end{equation} with probability 1, on the event that $G_* \in W_K$ and $\lambda^{5} \odot G_*$ does not contain $K$. Choose $h$ to maximize the RHS of~(\ref{eq dim bound}). Since the maximum is greater than 1, there is an $\epsilon > 0$ for which $$\dim (\partial T (K , G_* , h)) \geq 1 + \epsilon$$ with probability 1 on this event. Finally, let $\Omega$ be any connected component of $K^c$. By property (\ref{Whit-property} of the Whitney decomposition, for any open $V$ intersecting $\partial \Omega$, there is a tile $G_* \in W_K$ with $\lambda^{5} \odot G_* \subset V$, and the theorem follows from~(\ref{eq local dim}). $\fbox{\hphantom{} \vphantom{x}}$ \noindent{\bf Remark:} A planar domain $\Omega$ is called a {\bf John domain} if there is a base point $z_0 \in \Omega$ and a constant $C >0$ so that any point $x \in \Omega$ can be joined to $z_0$ by a curve $\gamma_x \subset \Omega$ so that $ \mbox{\rm dist}(z, \partial \Omega) \geq C|x-z| $ for any $z \in \gamma_x$. John domain were introduced by Fritz John in 1961, and some basic facts about them can be found in N{\"a}kki and V{\"a}is{\"a}l{\"a} (1994). With the notation of the above proof, if $G_* \in W_K$ is contained in a component $\Omega$ of $K^c$, choose for every tile $G \neq G_*$ in the tree $T (K , G_* , h)$, a Whitney chain leading to $G$ from its unique ancestor in the previous generation of the tree. For each tile $G'$ in this chain, there is an open disk containing it which is contained in $\lambda \odot G'$ (by Property \ref{inrad} of the tiling). The union of all these open disks as $G'$ runs over the chosen Whitney chain for $G$ and $G$ runs over $T (K , G_* , h)$, is a John domain $\Omega_{\rm J}$ satisfying $\dim (\partial \Omega \cap \partial \Omega_{\rm J}) \geq 1 + \epsilon$. \section{The traveling salesman theorem} \label{sec 4} Given a set $E$ in the plane and another bounded plane set $S$, we define $$ \beta_E (S) = (\mbox{\rm diam} (S))^{-1} \inf_{ L \in \cal L} \sup_{z \in E \cap S} \mbox{\rm dist} (z, L),$$ where $\cal L$ is the set of all lines $L$ intersecting $S$. \begin{th}[Jones 1990] \label{th jones} If $E \subset \mathop{\raise .45ex\hbox{${\bf\scriptstyle{|}}$ $ then the length of the shortest connected curve $\Gamma$ containing $E$ is bounded between (universal) constant multiples of $$\mbox{\rm diam}(E) + \sum_Q \beta_E (3 \odot Q)^2 \mbox{\rm diam} (Q) , $$ where the sum is over all dyadic squares in the plane and $3 \odot Q$ is the union of a 3 by 3 grid of congruent squares with $Q$ as the central square. \end{th} A simpler proof of this Theorem, and an extension to higher dimensions, are given in Okikiolu (1992). The theorem easily implies that the length $|\Gamma|$ of any curve $\Gamma$ which passes within $r$ of every point of $E$ satisfies \begin{equation} \label{eq jones} \mbox{\rm diam}(E) + \sum_{\mbox{\rm diam}(Q) \geq r} \beta_E (3 \odot Q)^2 \mbox{\rm diam} (Q) \leq c_3 |\Gamma| , \end{equation} where the sum is over all dyadic squares in the plane with diameter at least $r$. For every set $S$, there is a dyadic square $Q$ of side length at most $2 \mbox{\rm diam} (S)$ for which $S \subset 3 \odot Q$. Picking $S = \lambda^{5} \odot G$ for some tile $G$ and $Q$ accordingly, we get $$\beta_E (\lambda^{5} \odot G)^2 \mbox{\rm diam} (G) \leq 9 |\lambda|^{-10} \beta_E (3 \odot Q)^2 \mbox{\rm diam} (Q) $$ and since each expanded square $3 \odot Q$ contains a bounded number of expanded tiles $\lambda^{5} \odot G$ for tiles $G$ with $\sqrt{2} |\lambda|^{5} \mbox{\rm diam} (G) \geq \mbox{\rm diam} (Q)$, it follows that the length of any curve passing within $r$ of every point of $E$ satisfies \begin{equation} \label{eq jones2} |\Gamma| \geq c_4 \left ( \mbox{\rm diam}(E) + \sum_{\mbox{\rm diam}(G) \geq r} \beta_E (\lambda^{5} \odot G)^2 \mbox{\rm diam} (G) \right ) . \end{equation} We require the following corollary, which uses an idea from Bishop and Jones (1994). \begin{cor} \label{cor BJ} Let $\gamma$ be a Jordan curve with length denoted $|\gamma|$ and let ${\cal{C}}$ be a collection of Whitney tiles of index $n$. Let $U$ denote $\bigcup_{D \in {\cal{C}}} D$ and suppose that $\gamma \cup U$ is connected. Then there is a constant $c_5$ such that the cardinality of ${\cal{C}}$ is at least $$c_5 |\lambda|^n \left ( - |\gamma| + \sum_{G' \in \Xi ({\cal{C}})} \beta_U (\lambda^{5} \odot G')^2 \mbox{\rm diam} (G') \right ) \; , $$ where $\Xi ({\cal{C}})$ is the collection of tiles $G'$ of index at most $n$ for which $\lambda^{5} \odot G'$ intersects $U$. \end{cor} \noindent{\sc Proof:} Let ${\cal{C}}_\circ$ be the collection of circles of radius $r := |\lambda|^{-n} $ centered at points $\mbox{\rm center} (D)$ for $D \in {\cal{C}}$. Since neighboring tiles in ${\cal{C}}$ give rise to intersecting circles in ${\cal{C}}_{\circ}$, we see that $\Gamma := \gamma \cup \bigcup_{\Theta \in {\cal{C}}_{\circ}} \Theta $ is connected and passes within $r$ of every point of $\gamma \cup U$. Furthermore, any connected finite union of closed curves is a closed curve, and hence $\Gamma$ is a curve of length at most $|\gamma| + 2 \pi \# ({\cal{C}}) |\lambda|^{-n} $. Combining this with~(\ref{eq jones2}) shows that $$\# ({\cal{C}}) \geq {|\lambda|^n \over 2 \pi } \, \Big ( c_4 \sum_{\mbox{\rm diam} (G') \geq r} \beta_U (\lambda^{5} \odot G')^2 \mbox{\rm diam} (G') - |\gamma| \Big ) .$$ Since all tiles in $\Xi ({\cal{C}})$ have diameter at least $r$, this proves the lemma. $\fbox{\hphantom{} \vphantom{x}}$ \section{ Expected offspring in the Whitney tree} \label{sec 5} \noindent{\sc Proof of Lemma}~\ref{lem main}: Fix $G_*$ and $h$ as in the statement of the lemma and let $G$ be any tile in $T_n$. Let $\gamma$ be the circle separating $|\lambda|^4 \odot G$ from $\partial (|\lambda|^5 \odot G)$, which was used in Lemma \ref{lem wiggly -> beta}. Let ${\cal{C}}$ be the collection of tiles $D \in W_K^G$ of index $||G|| + h$ intersecting the disk bounded by $\gamma$. The union of all tiles in ${\cal{C}}$ is the set $U= U (G,h)$ defined before Lemma \ref{lem wiggly -> beta}. We want to show that the expected cardinality of $ T_{n+1} (G)$ is large. Since the cardinality of $ T_{n+1}(G)$ is at least $|\lambda|^{-14}$ times the cardinality of $\widetilde T_{n+1}(G)$ by (\ref{eq:prune}), and $\widetilde T_{n+1} (G) $ is a superset of ${\cal{C}}$, it suffices to show that $${\bf{E}} (\# {\cal{C}} \| {\cal{F}}_G) \geq c_1' h |\lambda|^h - c_2' |\lambda|^h \, .$$ To do this, we will apply Corollary~\ref{cor BJ} to ${\cal{C}}$, so that the set $U$ defined in that corollary is the same as $U(G,h)$ defined above. We will be able to bound from below the summands in Corollary~\ref{cor BJ} for most, but not all, ``intermediate-sized'' tiles $G'$. Pick an integer $a_3>1$ so that $|\lambda|^{3-a_3} < d_0$, where $d_0$ is the minimal distance between nonadjacent tiles in ${\cal{D}}_0$. Let $\widetilde \gamma$ be a circle concentric with $\gamma$, with a smaller radius: $\mbox{\rm rad}(\widetilde \gamma)=\mbox{\rm rad}(\gamma)-|\lambda|^{5-a_3}$ (see Figure \ref{GammaCir}). \begin{figure} \centerline{ \psfig{figure=GammaCir.ps,height=2.5in} } \caption{ \label{GammaCir} The circles $\gamma$ and $\tilde \gamma$ } \end{figure} For $a_3< j < h $, let $W_K^G (j)$ be the set of tiles $G' \in W_K^G$ such that $ ||G'||=||G|| + j$ and $\lambda^5 \odot G'$ intersects the disk bounded by $\widetilde \gamma$ . For any such tile $G'$ the blow-up $\lambda^5 \odot G'$ is contained in the disk bounded by $\gamma$. Fix any tile $G' \in W_K^G(j)$ with $a_3 < j \leq ||G|| + h - a_2$, where $a_2$ was specified in Lemma~\ref{lem wiggly -> beta}. That lemma implies \begin{eqnarray} && {\bf{P}} \left ( \beta_U (\lambda^{5} \odot G') \geq {\eta_0 \over 4 } \, \Big| \, {\cal{F}}_G \mbox {\rm and } W_K^G(j) \right ) \label{eq asdf} \\[2ex] & \geq & {\bf{P}} \Big[ \bigcap_{G''} \Big( K \, \mbox{ \rm $\eta_0$-surrounds } G'' \, \mbox { \rm \underbar{or} } \; K \cap \! \mbox{\rm core}(G'',\eta_0) = \emptyset \Big) \, \Big| \, {\cal{F}}_G \mbox {\rm and } W_K^G(j) \Big] \, \nonumber , \end{eqnarray} where the intersection is over all tiles $G'' \subset \lambda^{5} \odot G'$ such that $||G''|| = ||G'||$. The set of such tiles $G''$ for a fixed $G'$ has cardinality $|\lambda|^{10}$. Enumerating these and multiplying conditional probabilities using the hypotheses of Lemma~\ref{lem main} (since the $\sigma$-fields ${\cal{F}}_G$ and $\sigma(W_K^G(j))$ are contained in $\sigma(K \setminus G''^{\circ})$ ) gives a lower bound of $c_0^{|\lambda|^{10}} $ for~(\ref{eq asdf}), and implies that $$ {\bf{E}} \Big( \beta_U (\lambda^{5} \odot G')^2 \, \Big| \, {\cal{F}}_G \mbox{\rm and } W_G^K (j) \Big) \geq \left ( {\eta_0 \over 4 } \right )^2 c_0^{|\lambda|^{10}} = c_6>0 \, . $$ (This is the definition of $c_6$). Since $\gamma$ is outside $\lambda^4 \odot G$, the distance from $\gamma$ to $\lambda^3 \odot G$ is at least $|\lambda|^3 d_0 \cdot \mbox{\rm diam}(G)$, by Property \ref{min-dist} of the tiling. Therefore the distance from $\widetilde \gamma$ to $ \lambda^3 \odot G$ is at least $ (|\lambda|^3 d_0 - \lambda^{5-a_3} )\cdot \mbox{\rm diam}(G)$, which is greater than $ d_0 \cdot \mbox{\rm diam}(G) \,$ by the choice of $a_3$. By Lemma~\ref{lem 3}, the union of $\widetilde \gamma$ with all the tiles in $W_K^G (j)$ is a connected set. Since it intersects both $\lambda^3 \odot G$ and $\widetilde \gamma$, it follows that the cardinality of $W_K^G (j)$ is at least $d_0 |\lambda|^j$. Thus for each $j \in (a_3, h-a_2]$ we have $${\bf{E}} \Big( \sum_{G' \in W_K^G (j)} \beta_U (\lambda^{5} \odot G')^2 \, \Big| \, {\cal{F}}_G \mbox{\rm and } W_G^K (j) \Big) \geq c_6 d_0 |\lambda|^j . $$ By Corollary~\ref{cor BJ}, $${\bf{E}} (\# {\cal{C}} \| {\cal{F}}_G) \geq c_5 |\lambda|^{(n+1)h+||G_*||} \left ( -|\gamma| + \sum_{j = a_3+1}^{h - a_2} |\lambda|^{-nh-||G_*|| - j} c_6 d_0 |\lambda|^j \right ) . $$ Summing gives $${\bf{E}} ( \# {\cal{C}} \| {\cal{F}}_G) \geq c_5 \left ( c_6 d_0 (h - a_2-a_3) |\lambda|^h - |\gamma| \cdot |\lambda| ^{(n+1)h + ||G_*||} \right )$$ which proves the lemma since $|\gamma| = 2 \pi |\lambda|^{4 - nh - ||G_*||}$. $\fbox{\hphantom{} \vphantom{x}}$ \section{The Brownian frontier: Proof of Theorem 1.1} \label{sec 6} Let ${\bf{P}}_x$ denote the law of a planar Brownian motion $\{B(t) \}_{t \geq 0}$ started at $x$. We use ${\bf{P}}_0$ unless indicated explicitly otherwise. Let $\tau_{\exp}$ be a positive random variable, independent of the Brownian motion, which is exponential of mean 1 (i.e., its density is $e^{-t}$). We will verify~(\ref{eq gosper-wiggly}) for $K = B[0,\tau_{\exp}]$. by Brownian scaling, this will imply the first assertion of Theorem~\ref{th main}. \noindent{\bf Notation: } for any compact planar set $S$, denote by $\tau_S = \min \{ t \geq 0 \, : \, B(t) \in S \}$ the first hitting time of $S$, which is almost surely finite if $S$ has positive logarithmic capacity. Given $\eta \in (0,1/10)$, let $J_0$ be a rectifiable closed Jordan curve, which is the exterior boundary of a topological annulus separating $\mbox{\rm core}(G_0, \eta)$ from $\partial G_0$. (Here the constant $1/10$ can be replaced by any constant smaller than the inradius of $G_0$, and the existence of $J_0$ is guaranteed by Property \ref{annulus} of the tiling.) For the rest of this section, consider a homothetic image $G=z_{\rm cen}+r G_0$ of the Gosper Island $G_0$, with $r \in (0, 1)$ and $z_{\rm cen}$ in the plane. Also, denote by $J=z_{\rm cen}+r J_0$ the image of $J_0$ in $G$. We must obtain estimates which are uniform in the location and scale of $G$, as well as in the structure of the Brownian range outside $G$. \begin{lem} \label{lem:fastloop} For every $x \in \mbox{{\rm core}$(G,\eta)$ } $, \begin{equation} \label{eq:loop} {\bf{P}}_x \Big( B[0,\tau_J] \; \, \mbox{\rm $\eta$-surrounds } G \Big) \geq c_7(\eta) >0 \, . \end{equation} Furthermore, there exists $c_8(\eta)>0$ such that \begin{equation} \label{eq:fast} {\bf{P}}_x \Big( B[0,\tau_J] \; \mbox{ \rm $\eta$-surrounds } G \: \mbox{ \rm and } \tau_J < \tau_{\exp} < \tau_{_{ \partial G}} \Big) \geq c_8(\eta) {\bf{P}}_x(\tau_{\exp} < \tau_{_{ \partial G}}) \, . \end{equation} \end{lem} \noindent{\sc Proof:} The first estimate is immediate for $G_0$, and the general case follows by scaling. For the second, observe that by Brownian scaling, $\; \inf_{y \in J} {\bf{P}}_y (\tau_{_{ \partial G}} > \mbox{\rm diam}(G)^2) \: $ is a positive constant depending only on $J_0$, hence only on $\eta$. Also, clearly ${\bf{P}}(\tau_J < 1) >1/2$ and therefore $$ {\bf{P}}(\tau_J < \tau_{\exp} < \tau_J+ \mbox{\rm diam}(G)^2) > {e^{-2} \over 2} \mbox{\rm diam}(G)^2 \, , \; \mbox{ since } \, \mbox{\rm diam}(G) <1. $$ Applying (\ref{eq:loop}), lack of memory of exponential variables, and the strong Markov property of Brownian motion at the stopping time $\tau_J$, then shows that the left-hand side of (\ref{eq:fast}) is at least a constant multiple of $\mbox{\rm diam} (G)^2$. On the other hand, for any $x \in G$ we have ${\bf{P}}_x(\tau_{\exp} < \tau_{_{ \partial G}}) \leq {\bf{E}}_x ( \tau_{_{ \partial G}}) \leq \mbox{\rm diam}(G)^2$. This completes the proof. $\fbox{\hphantom{} \vphantom{x}}$ \begin{lem} \label{lem 3 parts} There exists $c_9 (\eta) > 0$ such that for any tile $G$, for any $x \in \mbox{{\rm core}$(G,\eta)$ } $ and any $A \subset \partial G$, $$ {\bf{P}}_x \Big(B[0,\tau_{_{ \partial G}}] \; \, \mbox{ \rm $\eta$-surrounds } G \; \mbox{\rm and } B(\tau_{_{ \partial G}}) \in A \Big) \geq c_9(\eta) {\bf{P}}_x (B(\tau_{_{ \partial G}}) \in A) . $$ \end{lem} \noindent{\sc Proof:} Recall that $J$ is a Jordan curve of finite length separating $\mbox{{\rm core}$(G,\eta)$ } $ from $\partial G$. The Harnack principle (see, e.g.,~Bass (1995, Theorem~1.20)) implies that there is a constant \newline $c_{10} = c_{10} (J_0)$ such that for any $y,z \in J$ and for any $A \subset \partial G$, \begin{equation} \label{eq:harnack} {\bf{P}}_y (B(\tau_{_{ \partial G}}) \in A) \geq c_{10} {\bf{P}}_z (B(\tau_{_{ \partial G}}) \in A) . \end{equation} Therefore for any $x \in \mbox{{\rm core}$(G,\eta)$ } $, \begin{eqnarray} &{\bf{P}}_x ( B[0,\tau_J] \; \, \mbox{ \rm $\eta$-surrounds } G \; \mbox{\rm and } B(\tau_{_{ \partial G}}) \in A ) & \nonumber \\[3ex] &= {\bf{E}}_x \Big( \mbox{\bf 1}_{\{ B[0,\tau_J] \; \mbox{ \rm $\eta$-surrounds } G \}} \cdot {\bf{P}}_{\! B(\tau_J)} [B(\tau_{_{ \partial G}}) \in A ] \Big) & \label{eq:one-ave} \end{eqnarray} Applying the Harnack inequality (\ref{eq:harnack}) with $y=B(\tau_J)$ and then invoking the estimate (\ref{eq:loop}) from the previous lemma, we find that the expression (\ref{eq:one-ave}) is at least $c_7(\eta) c_{10} {\bf{P}}_z (B(\tau_{_{ \partial G}}) \in A)$, for any $z \in J$. Finally, taking $z=B(\tau_J)$ and averaging with respect to ${\bf{P}}_x$ using the strong Markov property gives $$ {\bf{P}}_x ( B[0,\tau_J] \; \, \mbox{ \rm $\eta$-surrounds } G \mbox{\rm and } B(\tau_{_{ \partial G}}) \in A ) \geq c_7(\eta) c_{10} {\bf{P}}_x (B(\tau_{_{ \partial G}}) \in A) \, , $$ for any Borel set $A \subset \partial G$. This proves the lemma. $\fbox{\hphantom{} \vphantom{x}}$ \begin{figure} \centerline{ \psfig{figure=B1B2B3.ps,height=2.5in} } \caption{ \label{B1B2B3} The partition of the Brownian trajectory } \end{figure} Given $\eta>0$, we abbreviate $\tau_{_{\rm C}}= \tau_{\mbox{\rm core}(G, \eta)}$ and partition the Brownian trajectory into three pieces: \begin{enumerate} \item Until the first time $\tau_{_{\rm C}}$ that the path visits $\mbox{\rm core}(G, \eta)$. \newline Formally, define $B^{(1)}(t) = B(t \wedge \tau_{_{\rm C}})$ for $t \geq 0$, where $t \wedge s$ is shorthand for $\min \{t,s\}$. \item From time $\tau_{_{\rm C}}$ until the next visit to $\partial G$, denoted $\tau_{_{{\rm C}, \partial G}}= \min \{ t \geq \tau_c \, : \, B(t) \in \partial G \}$. \newline Define $B^{(2)}(t) = B((t+\tau_{_{\rm C}}) \wedge \tau_{_{{\rm C}, \partial G}})$ for $t \geq 0$. \item After time $\tau_{_{{\rm C}, \partial G}}$. \newline Denote $B^{(3)}(t)=B(t+\tau_{_{{\rm C}, \partial G}})$ for $t \geq 0$. \end{enumerate} The idea now is that $B^{(1)}$ and $B^{(3)}$ determine the Brownian range outside $G^\circ$, and $B^{(2)}$ has a substantial chance of $\eta$-surrounding $G$, even when we condition on its endpoints. However, we still have to take the exponential killing into account. Define the random variable $$ I = \left\{ \begin{array}{lcr} 1 & \mbox{\rm if} & \tau_{\exp} < \tau_{_{\rm C}} \\ 2 & \mbox{\rm if} & \tau_{_{\rm C}} \leq \tau_{\exp} < \tau_{_{{\rm C}, \partial G}} \\ 3 & \mbox{\rm if} & \tau_{_{{\rm C}, \partial G}} \leq \tau_{\exp} \end{array} \right. \, $$ that indicates in which part of the motion the exponential killing occurred. Finally, define $$ \widetilde \tau_{\exp} = \left\{ \begin{array}{lcr} \tau_{\exp} & \mbox{\rm if} & I=1 \\ \tau_{_{\rm C}} & \mbox{\rm if} & I=2 \\ \tau_{\exp}-\tau_{_{{\rm C}, \partial G}} & \mbox{\rm if} & I=3 \end{array} \right. $$ \begin{pr} \label{pr:brown} For any $\eta \in (0, 1/10)$ there is a constant $c_0= c_0(\eta) > 0$ such that for all homothetic images $G=z_{\rm cen} +r G_0$ of $G_0$ with $r \in (0, 1)$ and $z_{\rm cen}$ in the plane: \begin{eqnarray} \label{eq gosper-pf} {\bf{P}} \Big ( B[0, \tau_{\exp}] \, \mbox{ \rm $\eta$-surrounds } G \, \mbox { \rm \underbar{or} } \; B[0, \tau_{\exp}] \cap \mbox{{\rm core}$(G,\eta)$ } = \emptyset \, \, \Big| \, \, {\cal{A}}_G \Big ) > c_0 \; \; \: & \; & \; \end{eqnarray} where the conditioning is on the $\sigma$-field ${\cal{A}}_G$ generated by $I, \, B^{(1)}, \, B^{(3)} \mbox{\bf 1}_{\{I=3 \} } \,$ and $\, \widetilde \tau_{\exp}$. \end{pr} \noindent{\sc Proof:} \, On the event $\{I=1\}$, the set $B[0, \tau_{\exp}]$ is disjoint from $\mbox{{\rm core}$(G,\eta)$ } $. \newline To handle the case $\{I=2\}$, we use the strong Markov property at time $\tau_{_{\rm C}}$ and apply the estimate (\ref{eq:fast}) to $B^{(2)}$. Denoting $\tau_{_{{\rm C}, J}}= \min \{ t \geq \tau_c \, : \, B(t) \in \partial G \}$, this gives \begin{eqnarray*} & {\bf{P}}_0 \Big( B[\tau_{_{\rm C}},\tau_{_{{\rm C}, J}}] \; \mbox{ \rm $\eta$-surrounds } G \: \mbox{ \rm and } \tau_{_{{\rm C}, J}} < \tau_{\exp} < \tau_{_{{\rm C}, \partial G}} \, \Big| \, I \geq 2 ; \, B^{(1)} \Big) & \\[3ex] & \geq c_8(\eta) {\bf{P}}_0 \Big( \tau_{_{\rm C}} < \tau_{\exp} < \tau_{_{{\rm C}, \partial G}}) \, \Big| \, I \geq 2 ; \, B^{(1)} \Big) \, . & \end{eqnarray*} This proves (\ref{eq gosper-pf}) on the event $I=2$. Only the case $I=3$ remains. By using the strong Markov property at time $\tau_{_{\rm C}}$ and applying Lemma \ref{lem 3 parts} to $B^{(2)}$, we see that for any $A \subset \partial G$, $$ {\bf{P}}_0 \Big(B[\tau_C,\tau_{_{{\rm C}, \partial G}}] \; \, \mbox{ \rm $\eta$-surrounds } G \; \mbox{\rm and } B(\tau_{_{{\rm C}, \partial G}}) \in A \, \Big| \, B^{(1)} \Big) \geq c_9(\eta) {\bf{P}}_0 \Big(B(\tau_{_{{\rm C}, \partial G}}) \in A \, \Big| \, B^{(1)} \Big) \, . $$ In other words, $$ {\bf{P}}_x \Big( B[\tau_C,\tau_{_{{\rm C}, \partial G}}] \; \, \mbox{ \rm $\eta$-surrounds } G \, \Big| \, B^{(1)}, \, B(\tau_{_{{\rm C}, \partial G}}) \Big) \geq c_9(\eta) \, . $$ An application of the strong Markov property at time $\tau_{_{{\rm C}, \partial G}}$ shows that this lower bound is still valid if we insert an additional conditioning on $I \geq 2$ and on $B^{(3)}$. \newline Finally, since ${\bf{P}}_0(I =3 \| I \geq 2) \geq e^{-1}/2$ and $\widetilde \tau_{\exp}$ is conditionally independent of $B[0,\tau_{_{{\rm C}, \partial G}}]$ given $I=3$, this completes the proof of the proposition. $\fbox{\hphantom{} \vphantom{x}}$ To obtain the uniformity in Theorem~\ref{th main}, we will need the following general observation. \begin{lem} \label{lem semicont} Let $\Gamma : [0,\infty) \rightarrow \mathop{\raise .45ex\hbox{${\bf\scriptstyle{|}}$$ be any continuous path and let $t > 0$. For any open disk $U$ intersecting $\mbox{\rm \small frontier}(\Gamma[0,t])$ such that $\Gamma (t) \notin \overline{U} $, there is a $\delta > 0$ such that for any $s \in [t , t + \delta ]$, we have \begin{equation} \label{eq:contain} U \cap \mbox{\rm \small frontier} ( \Gamma [0,t]) \, \supset \, U \cap \mbox{\rm \small frontier} (\Gamma [0,s]) \, \neq \emptyset \, . \end{equation} \end{lem} \noindent{\sc Proof:} By hypothesis $U$ intersects the unbounded component, $\Omega$. of $\Gamma [0,t]^c$, so there is a point $u\in U$ with and an unbounded curve starting from $u$ and contained in $\Omega$. Using the convexity of $U$, we can append to this curve a line-segment connecting $u$ to a nearest point $x$ on $\Gamma [0,t]$, and thus obtain an unbounded curve $\psi$ starting at $x$ and contained in $\Omega \cup \{x\}$. Choose $\delta>0$ small enough so that $\Gamma [t,t + \delta]$ is disjoint from the curve $\psi$. This gives the right-hand side of (\ref{eq:contain}) for $s \in [t , t + \delta ]$. If we also require that $\Gamma [t,t + \delta]$ is disjoint from $U$, then the left-hand side of (\ref{eq:contain}) follows from the general fact that $$ \mbox{\rm \small frontier} ( \Gamma [0,s]) \subset \mbox{\rm \small frontier} (\Gamma [0,t]) \cup \Gamma [t,s] \, . $$ $\fbox{\hphantom{} \vphantom{x}}$ \noindent{\sc Proof of Theorem~\ref{th main}:} $\:$ The random set $B[0, \tau_{\exp}] \setminus G^\circ$ is completely determined by the variables generating the $\sigma$-field ${\cal{A}}_G$ defined in Proposition \ref{pr:brown}, so the proposition implies that $K=B[0, \tau_{\exp}]$ satisfies the hypothesis (\ref{eq gosper-wiggly}) of Theorem \ref{th any K}. Since $B[0,\tau_{\exp}]$ has the same distribution as $\sqrt{\tau_{\exp}} \cdot B[0,1]$, this establishes the first assertion of Theorem~\ref{th main}. For $t>0$, Let $A_t$ be the event that $\dim (\mbox{\rm \small frontier} (B[0,s]) \cap V) \geq 1 + \epsilon$ simultaneously for all open disks $V$ that intersect $\mbox{\rm \small frontier}(B[0,t])$ and have rational centers and radii. Theorem \ref{th any K} and Proposition \ref{pr:brown} give ${\bf{P}}_0(A_1)>0$, and we must show that ${ \displaystyle {\bf{P}}_0 ( \cap_{t>0} A_t ) = 1}$. Denote by $A_{\rm I\!\!\!Q}= \cap_{s \in {\rm I\!\!\!Q}_+} A_s$ the intersection over all positive rational times. Now Brownian scaling and countable additivity imply that ${\bf{P}}_0 (A_{\rm I\!\!\!Q})=1$, so it suffices to prove that $A_{\rm I\!\!\!Q} \subset A_t$ for all $t>0$. Fix $t>0$ and an open disk $V$ that intersects $\mbox{\rm \small frontier}(B[0,t])$. Since $\mbox{\rm \small frontier} (B[0,t])$ is connected, it must intersect some (random) open disk $U = U(V,t)$ with rational center and radius such that $U \subset V$ and $B(t) \notin \overline{U}$. By the previous lemma, there is a rational $s$ such that $$ \mbox{\rm \small frontier} (B[0,t]) \cap U \, \supset \, \mbox{\rm \small frontier} (B[0,s]) \cap U \neq \emptyset \, . $$ This implies that $A_{\rm I\!\!\!Q} \subset A_t$, and completes the proof of the theorem. $\fbox{\hphantom{} \vphantom{x}}$ Finally, we consider the {\bf planar Brownian bridge} $B_{\rm br}$, which may be defined either by conditioning the Brownian path to return to the origin, or by $B_{\rm br}(t) = B(t)-tB(1)$ for $t \in [0,1]$. For every $t<1$, the restrictions $B_{\rm br}|_{[0,t]}$ and $B|_{[0,t]}$ have mutually absolutely continuous laws (these laws are measures on the space of continuous maps from $[0,t]$ to the plane.) Therefore by Theorem \ref{th main}, for every {\em fixed} $t \in (0,1)$, \begin{equation} \label{eq:bridge1} \dim \Big(\mbox{\rm \small frontier}(B_{\rm br}[0,t]) \Big) \, \geq 1 + \epsilon \; \, \mbox{ a.s.} \end{equation} Consider a sequence of annuli $\{A_n\}$ of modulus $2^{-n}$ around the origin. The probability that $B_{\rm br}$ surrounds the origin in $A_n$ is bounded away from 0, so the Blumenthal 0--1 law implies that with probability 1, there is some rational $t<1$ such that $\mbox{\rm \small frontier}(B_{\rm br}[0,1]) = \mbox{\rm \small frontier} (B_{\rm br}[0,t]) $ (see Burdzy and Lawler (1990)). Thus by (\ref{eq:bridge1}), with probability 1, $$ \dim \Big(\mbox{\rm \small frontier}(B_{\rm br}[0,1]) \Big) \, \geq \inf_{t \in {\rm I\!\!\!Q} \cap (0,1)} \, \dim \Big(\mbox{\rm \small frontier}(B_{\rm br}[0,t]) \Big) \geq \, 1 + \epsilon \,. $$ \section{Concluding remarks} It can be shown (Krzysztof Burdzy, personal communication) that $\dim (\mbox{\rm \small frontier}(B[0,1]))$ is almost surely constant; this fact is not required for the arguments in this paper. The conjecture that the Brownian frontier has dimension $4/3$ is related to well-known conjectures concerning self-avoiding random walks, which in turn are a model for long polymer chains. In that context, the exponent $4/3$ first appeared in the non-rigorous considerations of Flory (1949); see also de Gennes (1991). Theorem \ref{th any K} is stated for general random sets, rather than just Brownian motion, in view of potential applications to the ranges and level-sets of other stochastic processes. Besides the range of Brownian motion, another natural random set that satisfies the hypothesis of Theorem \ref{th any K} is the support of {\bf super-Brownian motion}, i.e. the intersection of all closed sets that are assigned full measure by this measure-valued diffusion throughout its lifetime. (For the definitions see, e.g., Dawson, Iscoe, and Perkins (1989).) Equivalently, this random set may be characterized as the set of points ever visited by the path-valued process constructed by Le-Gall (1993). (This process is often referred to as ``The Brownian snake''.). We are grateful to Steve Evans for enlightening discussions of super-Brownian motion. To allow for further applications, we state below a variant of Theorem \ref{th any K} which obtains the same conclusions under weaker hypotheses on the random set $K$. We omit the proof, which requires the estimates obtained by Pemantle (1994) for the probability that a Wiener sausage covers a straight line segment. For any set $S \subset \mathop{\raise .45ex\hbox{${\bf\scriptstyle{|}}$$ and any $\epsilon > 0$, let $S^\epsilon$ denote the set $\{ x : |x-y| \leq \epsilon \mbox{ for some } y \in S \}$. Say that $K$ {\em is $\eta , \delta$-flat} inside $S$ if there is some line segment $\ell$ of length $\eta \mbox{\rm diam} (S)$ covered by $K^{\delta \mbox{\rm diam} (S)}$, having $\ell^{\eta \mbox{\rm diam} (S)}$ inside $G$ with $\ell$ not topologically surrounded by $K \cap G \cap (\ell^{\delta \mbox{\rm diam} (S)})^c$. \begin{th} \label{th any K2} Let $G_0$ be the Gosper Island, and let $K$ be a random compact connected subset of the plane. Suppose that for some $\delta_0 > 0$, the following hypothesis on $K$ is satisfied, where the supremum is over $r \in (0,1)$ and ${\bf x}$ in the plane. \begin{equation} \label{eq sup condition} \sup_{G = {\bf x} + r G_0} \mbox {\rm ess sup}\; {\bf{P}} \left [ K \mbox{ is } \eta_0 , \delta_0 \mbox{-flat inside } G \| K \cap G \neq \emptyset , \sigma (K \cap G^c) \right ] < 1 . \end{equation} Then there is an $\epsilon > 0$ for which $\; \dim (\mbox {\rm \small frontier}(K)) \geq 1 + \epsilon \, $ with probability 1. \end{th} \noindent{\sc Remark:} The intuition behind the two-part definition of $\eta , \delta$-flatness is that for $\mbox {\rm \small frontier}(K)$ to be close to straight (thus for $K$ to be flat), $K$ itself must nearly cover a line segment and this must happen somewhere that is not completely encircled by $K$. For the special case when $K$ is the range of planar Brownian motion, it seems likely that methods directly adapted to this case will yield better estimates for $\dim (\mbox {\rm \small frontier}(K))$ than those obtainable by our methods. Indeed, Gregory Lawler has informed us that immediately after he learned of our Theorem \ref{th main} (but without seeing its proof), he proved (using completely different methods) that the dimension of the Brownian frontier can be expressed in terms of the ``double disconnection exponent'' of Brownian motion. This allowed Lawler to deduce that $\dim (\mbox{\rm \small frontier}(B[0,1])) > 1.01$ a.s., by invoking recent estimates of Werner (1994) on disconnection exponents. We refer the reader to Lawler's forthcoming paper for this and several other striking results on the Brownian frontier. Finally, we note an application to simple random walk on the square lattice $Z\!\!\!Z^2$. Given a subset $S$ of $Z\!\!\!Z^2$, say that a lattice point $x \in S$ is on the {\bf outer boundary} of $S$ if $x$ is adjacent to some point in the unbounded component of $Z\!\!\!Z^2 \setminus S.$ We remark that using the strong approximation results of Auer (1990) and our construction of the Whitney tree in Section 3, it is easy to derive the following. \begin{cor} \label{cor:srw} Let $\{S(k)\}$ denote simple random walk on $Z\!\!\!Z^2$, and let $\epsilon>0$ be as in Theorem \ref{th main}. Then for every $\epsilon_1 < \epsilon$ we have $$ \lim_{n \to \infty} {\bf{P}}\{ \mbox{There are more than } \: n^{(1+\epsilon_1)/2} \, \mbox{ points on the outer boundary of } \: S[0,n]. \} = 1 $$ \end{cor} \renewcommand{\baselinestretch}{1.0}
\section{Introduction} The main motivation for the models of CP violation has been to explain the $\epsilon'/\epsilon $ parameter for neutral kaon decay simultaneously preserving the vanishing smallness of the electric dipole moment of the neutron. The idea that CP violation can be accomodated in extensions of the Higgs sector was put forward by Weinberg \cite{Wein1} in his three Higgs doublet model, and T.D. Lee \cite{Lee1} in his two Higgs doublet model (2HDM) with spontaneous CP violation. In these models the Yukawa couplings have to be constrained to suppress flavor changing neutral currents. A natural condition is to require that either $\phi_1$ couples to the up quarks and $\phi_2$ to the down quarks or $\phi_1$ couples to all the quarks and $\phi_2$ does not couple at all. A discrete symmetry $\phi_1 \longrightarrow -\phi_1$ is imposed on the Higgs potential to ensure suppression. In majority opinion the minimal standard model suffers from inadequate CP violation coming from the CKM matrix to be considered a viable model for baryogenesis at the electroweak scale. So the 2HDM with with CP violation in the Higgs sector has attracted attention for the past several years in the context of electroweak baryogenesis. McLerran-Shaposhnikov-Turok and Voloshin \cite{MacL} had shown that in this model, the triangle diagram with one gauge boson leg and two scalar legs and top quark in the loop contributes a correction to the effective action $\sim {\cal O} N_{cs}$, where $\cal O$ is an operator that can be found out by explicit evaluation of the diagram and $N_{CS}$ is the Chern-Simons (CS) number. This correction biases the CS number and since $\bigtriangleup N_B = n_{f} \bigtriangleup N_{CS}$, nett baryon number results, with the sign determined by the sign of $<\cal O >$. Recently Yue-Liang-Wu \cite{Wu} has generalised the model of Glashow and Weinberg \cite{Wein2} and has proposed that the value of $\epsilon'/\epsilon $ and the limits on EDM can be well explained with less stringent conditions on the Yukawa couplings. The Higgs boson-fermion couplings are now much more general and no discrete symmetries are imposed on the Higgs potential. In this class of models fermions acquire mass from both $\phi_1$ and $\phi_2$. Consequently in addition to the triangle diagrams with $\phi_1$ on the external legs, diagrams with $\phi_2$ also on the external legs start contributing to the effective action. The MSTV proposal has been criticised on the grounds that it is a higher order adiabatic effect, of the order $(\langle \phi \rangle^T/T)^4$. In their considerations the relative phase between the two Higgs vacuum expectation values was treated constant throughout bubble evolution. Here we provide the details of the argument \cite{sbdanduay2} that the inclusion of the dynamics of the relative phase removes the adiabatic suppression, making it an effect of the order $(\langle \phi \rangle^T/T)^2$. We have also previously shown \cite{sbdanduay1} that the uncertainties of physical parameters of bubble formation \cite{Dineetal} can be circumvented if we consider a string induced phase transition \cite{Yajnik}. This is the scenario we shall be considering here as well. It may be noted that since according to \cite{sbdanduay1}, the string induced bubbles provide adiabatic conditions, all B-genesis mechanisms relying on such conditions are workable. In particular the mechanisms of Cohen and Kaplan \cite{Cohen2} and Cohen-Kaplan-Nelson \cite{Cohen3} can also proceed through string induced bubbles. We recapitulate here the corresponding bubble profile ansatz including the time evolution of the relative phase and calculate the resultant contribution to the effective action. Putting together the two sources of enhancement, viz., dynamics of the relative phase and additional triangle diagrams, we show that the mechanism has sufficient potential to give adequate baryogenesis. \section{The triangle diagram and correction to the effective action} It was first pointed out by Turok and Zadrony \cite{Turok} and then explicitly calculated by McLerran-Shaposhnikov-Turok and Voloshin \cite{MacL}, that for 2HDM a term biasing the Chern-Simons number with a CP odd chemical potential is contributed by the triangle diagram of figure 1. This contribution to the effctive action at finite temperature is \begin{equation} \Delta S={-7\over{4}}{\zeta(3)}{\left(m_{t}\over{\pi T}\right)^2} { g\over 16{{\pi }^2}} {1\over{{v_1}^2}} \times \int ( {\cal D}_{i}\phi_{1}^{\dagger}\sigma^{a}{\cal D}_{0}\phi_1 + {\cal D}_{0}\phi_{1}^\dagger \sigma^{a}{\cal D}_{i}\phi_1) \epsilon^{ijk} F_{jk}^{a}d^{4}x \end{equation} where $m_{t}$ is the mass of the top quark, $\zeta $ is the Riemann Zeta function, and $\sigma^{a}$ are the Pauli matrices. For homogeneous but time varying configurations of the Higgs fields and in the ${A_0}^{a}=0$ gauge, \begin{eqnarray} \Delta S_1 &=& {-i7\over{4}}{\zeta(3)}{\left(m_{t}\over{\pi T}\right)^2} {2\over{{v_1}^2}} \int dt[{\phi_1}^\dagger{\cal D}_{0}\phi_{1} - {\cal D}_{0}\phi_1^\dagger \phi_{1}] N_{CS}\\ & =& {\cal O}_{1} N_{CS} \end{eqnarray} Clearly, here the top quark acquires mass only from $\phi_1$ as demanded by the Glashow-Weinberg natural flavor conservation (NFC) criteria. In Wu's model, the FC criterion is satisfied under relaxed conditions on the Higgs-fermions couplings. In this case the general Yukawa interactions can be written as \begin{equation} L_Y = \bar{q_{iL}} {(\Gamma_{D}^{a})}_{ij} D_{jR} \phi_a +\bar{q_{iL}} {(\Gamma_{U}^{a})}_{ij} U_{jR} \bar{\phi_a} + \bar{l_{iL}} {(\Gamma_{E}^{a})}_{ij} E_{jR} \phi_a + h.c \end{equation} where $q_{i}$, $l_i$ and$\phi_a$ are $SU(2)_L$ doublet quarks, leptons and Higgs bosons, $U_i$, $D_i$, $E_i$ are $SU(2)_L$ singlets. $i = 1,2,...n $ is a generation label and $a = 1,2$ is a Higgs doublet label. ${\Gamma^a}_F$ $(F = U, D, E)$ are the Yukawa coupling matrices. According to Wu the Glashow-Weinberg criteria can be replaced by a theorem which states that the flavor conservation for the neutral currents is natural in the Higgs sector or equivalently, the matrices ${\Gamma^a}_F$ $(F = U, D, E)$ are diagonalizable simultaneously by a biunitary or biorthogonal transformation, if and only if the square $n \times n$ ${\Gamma^a}_F$ are represented in terms of the linear combinations of a complete set of $n \times n$ matrices $ ({\Omega^\alpha}_F, \alpha = 1,2,...n)$. \begin{equation} {\Gamma^a}_F = \sum_{\alpha} {g^F}_{a\alpha} {\Omega_F}^\alpha \end{equation} where, ${\Omega_F}^\alpha$ satisfy the following orthogonal condition ${\Omega_F}^\alpha {({\Omega_F}^\beta)}^\dagger = {L^\alpha}_F \delta_{\alpha\beta}$,\\ ${({\Omega_F}^\alpha)}^\dagger {\Omega_F}^\beta= {R^\alpha}_F \delta_{\alpha\beta}$ , with the normalization $\sum_\alpha {L^\alpha}_F = \sum_\alpha {R^\alpha}_F = 1$. This generalised Yukawa coupling prompts us to consider another set of triangle diagrams as shown in fig-2. The contribution at finite temperature from these diagrams to the effective action can similarly be calculated to be \begin{equation} \Delta S_2 = {-i7\over{4}}{\zeta(3)}{\left(1\over{\pi T}\right)^2} {{\Gamma}^{2}}^2 \int dt[{\phi_2}^\dagger({\cal D}_{0}\phi_{2}) - ({\cal D}_{0}\phi_2)^\dagger \phi_{2}] N_{CS} \end{equation} \begin{equation} \Delta S_3 = {-i7\over{4}}{\zeta(3)}{\left(1\over{\pi T}\right)^2} {\Gamma}^{1}{\Gamma}^{2}\int dt[{\phi_1}^\dagger({\cal D}_{0}\phi_{2}) - ({\cal D}_{0}\phi_1)^\dagger \phi_{2}] N_{CS} \end{equation} where $\Gamma_{1(2)}$ is the Yukawa coupling when the fermions couple to $\phi_{1(2)}$. \subsection{The bubble profile} Now to find the bubble profile we use the following ansatze for the finite temperature vacuum expectation value of the Higgs fields \begin{eqnarray} {\phi_{1}}^0 &=& {\rho}_{1}(r,t) e^{-i\theta(t)} \\ {\phi_{2}}^0 &=& {\rho}_{2}(r,t) e^{i\omega(t)} \end{eqnarray} where, as pointed out by Cohen-Kaplan-Nelson \cite{Cohen1} we can fix the unitary gauge ensuring $\theta$ is the physical pseudo-scalar orthogonal to the Goldstone boson eaten by Z. This gauge fixing gives the relation between $\omega$ and $\theta$ to be, \begin{equation} \partial_\mu \omega = {(\rho_1 / \rho_2)}^2 \partial_\mu \theta \end{equation} We assume that the phase transition takes place when a combination of $\rho_{1}$ and $\rho_{2}$ becomes massless in a particular direction $\gamma $ in the ${\phi_{1}}^0 - {\phi_{2}}^0 $ plane. Hence, in the bubble profile we may take \begin{equation} {\rho}_1(r,t)=\rho(r,t)\cos{\gamma} \hskip 1true in \rho_2(r,t)= {\rho}(r,t)\sin{\gamma} \end{equation} with this parametrization and finite temperature corrections, the effective potential for the $\rho$, $\theta$ and $\omega$ is \begin{eqnarray} V_{T}(\rho,\theta,\omega)&=&{M_1}^{2}(T){\rho}^2 - ET{\rho}^{3} + {K_{1} \over {4}} {\rho}^4 \nonumber \\ &&+ {M_2}^{2}(T)(\cos{\xi} \cos{(\omega + \theta)} + {C_1}\sin{\xi}\sin{(\theta + \omega)} ){\rho}^2 \nonumber \\ &&+ {K_{2}\over {4} }{\rho}^4 (\cos^2{(\theta + \omega)} +{C_{1}} \sin^{2}{(\theta + \omega )}) \end{eqnarray} where, ${M_{1}}^2$ and ${M_{2}}^2$ are temperature corrected mass parameters and $K_{1}$ and $K_{2}$ are combinations of quartic coupling constants with small temperature dependent corrections. The constant $C_{1}$ is the ratio of ${\lambda _5}$ and $\lambda _{6}$ in the standard parameterisation of the 2HDM \cite{hhguide}, and $\xi$ is the phase of the neutral component of zero temperature vacuum expectation value of $\phi_2$. Rescaling ${\rho \longrightarrow ({2ET/{K_{1}}}){\rho}^{\prime}}, r \longrightarrow{{r^\prime}/{2ET\over {K_{1}}}}$ and $t$ $\longrightarrow{{t^\prime} /{2ET\over {K_{1}}}}$ omitting the primes the potential can be written as, \begin{eqnarray} V_{T}(\rho,\theta, \omega)&=&{\left( {2ET}\over {K_{1}}\right)}^{4} [{K_{1}\over{2}}(1+E_{1}){\rho}^{2} - {K_{1}\over {2}}{\rho}^{3}+ {K_{1}\over {4}} {\rho}^{4} \nonumber \\ &&+ {K_{2}\over{2}}(1+E_{2})(\cos{\xi} \cos(\omega + \theta)+ {C_1}\sin\xi \sin(\theta + \omega) ){\rho}^{2} \nonumber \\ &&+ {K_{2}\over {4} }{\rho}^4 (\cos^2(\theta + \omega) + {C_{1}} \sin^{2}(\theta + \omega ))] \end{eqnarray} where $E_{1} = \frac{M_{1}^{2}K_{1}}{2 E^{2} T^{2}}-1 $ and $E_{2} = \frac{M_{2}^{2}K_{1}}{2 E^{2} T^{2}}-1 $. The geometry of string induced bubbles is cylindrical. With this in mind, the time dependence of $\rho$ and $\theta$ can be found by solving the following equations, \begin{eqnarray} \lefteqn{ {\frac{{\partial}^{2}\rho}{\partial t^{2}}} - {\frac{{\partial}^{2}\rho}{\partial r^{2}}} -\frac{1}{r}\frac{\partial\rho}{\partial r} + \rho{\left( {\partial \theta}\over {\partial t} \right)}^2 - {3\over2}K_{1}{\rho}^2 + K_1{\rho}^3+ {K_{1}\over 2}(1+E_{1})\rho } \nonumber\\ && +K_{2}C_{1} \sin^{2}(\theta + \xi){{\rho}^3} + {K_{1}\over 2}(1+E_{2}) (\cos{\theta} + C_{1} \sin{\xi} \sin{(\xi + \theta}))\rho = 0 \end{eqnarray} \begin{eqnarray} \lefteqn{ \frac{\partial^2\theta}{\partial t^2} + \frac{K_2}{4} C\sin{2(\theta+\xi)}{\rho^2} } \nonumber\\ &&+\frac{K_1}{2}(1+E_2)(C \sin\xi\cos{(\theta + \xi)} -\sin(\theta+\omega-\xi)) = 0 \end{eqnarray} The time independent solution of $\rho$ can be found by imposing the boundary condition $\rho \longrightarrow 0$ as $r \longrightarrow \infty $ and ${{\partial \rho}\over{\partial r}} \longrightarrow 0$ as $ r \longrightarrow 0$. Subsequently, as the nontrivial minimum becomes favorable, the same solutions begin to evolve in time. The parameters used in the equations are, \noindent \centerline{\vbox{\noindent $K_{1}$, $K_{2}$ = 0.1 -- $0.001$, \\ $C_{1} = 1.1$, $C = C_{1}-1$, $E_{1} = -0.074$, $E_{2} = -0.07$,\\ $\xi = 0.2$, $E\sim 0.01$}} The time evolution of the bubble profile $\rho(r,t)$ has been reported earlier \cite{sbdanduay1}. In fig-3, we show the time evolution of the relative phase in the regions where $\rho$ has become nonzero. If the initial reference value is zero, it oscillates to reach the stationary value dictated by the 2HDM effective potential at the relevant temperature. We assume this value to be $O(1)$ since no natural reasons prevent it from being so. \subsection{Evaluation of the operator} Now we can use these solutions to evaluate the average value of the operators as, \begin{equation} {\cal O}_1 + {\cal O}_2 = 28 {\zeta(3)}{\left(1\over{\pi }\right)^2} { E^{2} \over {K_{1}^2}}A_{1} {\int {{\rho}^2}{\frac{\partial \theta}{\partial t}} dt} \end{equation} and, \begin{equation} {\cal O}_3 = 28 {\zeta(3)}{\left(1\over{\pi }\right)^2} { E^{2} \over {K_{1}^2}}A_{2} {\int {{\rho}^2}{\frac{\partial \theta}{\partial t}}{e^{i(\theta + \omega)} dt}} \end{equation} where $A_1 = {({\Gamma_1}\cos\gamma)}^2 + {({\Gamma_2}\sin\gamma)}^2$ and $A_2 =\Gamma_1\Gamma_2 sin2\gamma$. These parameters have to be determined from the phenomenology of the 2HDM which is yet far from being tested by current experiments. Knowing the solutions to eqn.s (14)-(15), we can estimate the integral in eqn. (16) to be ${\rho_{\infty}}^2\Delta\theta$ where $\Delta\theta$ is the nett change in the relative phase $\theta$ at any given point as the bubble wall sweeps past it, and $\rho_{\infty}=1$. There is an additional contribution from the transient part of $\theta$, which can also be calculated numerically, but is not significant. As for the term ${\cal O}_3$, its CP odd part has the magnitude $\rho_{\infty}^2\Delta(\sin\theta)$ from arguments already given. Putting in other known factors, we see that \begin{equation} {\cal O}_1 + {\cal O}_2 \simeq\ A_1\left(E/K_1 \right)^2 \Delta\theta \end{equation} and a similar contribution from ${\cal O}_3$. Recall that $E$ is the dimensionless cubic self-coupling induced by thermal loops, and $K_1$ involves the quartic self-couplings of the 2HDM. For naturalness we would like $\Delta\theta$ to be $O(1)$ but the remaining factor is numerically a small magnitude, perhaps between $1$ and $10^{-4}$. Note however that the effect is not suppressed by the physics of the process viz., the bubble wall dynamics. We emphasise again that this is the consequence of the dynamics of the relative phase of the 2HDM. \section{Estimation of the asymmetry} To estimate the baryon asymmetry we assume the presence of high temperature sphaleron processes inside the bubble wall. The rate of such transitions per unit time per unit volume is of the order $\sim \kappa{\alpha_{w}T}^4$, $\kappa \sim 1$ \cite{Amb}. The number of fermions created per unit time in the bubble wall is given by \begin{equation} B = \kappa {(\alpha _{w} T)}^{4} l S\times {\frac {1}{T}}{\frac {{\cal O}} {l}} \end{equation} where we have made use of a well established master formula \cite{BocShap} \cite{Cohen1}, and where $l$ and $S$ are the thickness and the surface area of the bubble wall respectively. From which we get the baryon to photon ratio to be \begin{eqnarray} \bigtriangleup \equiv {\frac {n_{B}}{s}}& \simeq& {\frac{1}{N_{eff}}} {(\alpha_{w})}^4 {\cal O } \nonumber \\ & \simeq & 10^{-8}\times \left( \frac{E}{K_1} \right)^2 \Delta\theta \end{eqnarray} where we have used $\alpha_w$ $\sim$ $10^{-\frac{3}{2}}$ and $N_{eff} \sim 100$. This answer easily accomodates the observed value of this number. It is worth emphasising the physics of this answer which is fairly robust against changes in the specific particle physics models. The thermal rate contributes $10^{-6}$ through $\alpha_w^4$, and another $10^{-2}$ is contributed by $N_{eff}$. The remaining smallness of the answer follows from smallness of the thermal vacuum value $\langle \phi \rangle^T$ $\sim ET/K_1$, which is small on the scale set by $T_c$. The appearance of this particular physical quantity has to do with our picture of the process occuring in bubble walls as the phase transition is yet in progress. \section{Conclusion} The MSTV proposal has been criticised \cite{Cohen1} on the grounds that the operator ${\cal O}$ is of the order of $(\phi/T)^4$. Since $\phi$, i.e., the temperature dependent vacuum value $\langle \phi \rangle^T$ is $\ll T$, the effect was thought to be unacceptably small. But with time variation of the relative phase allowed, we find the relevent operator is of the order of $(\phi/T)^2$. The effect is therefore not intrinsically suppressed. We also note that the same analysis could be fruitfully applied to the proposals of \cite{Cohen2} and \cite{Cohen3}. Secondly, we have shown that in principle there are more diagrams in the 2HDM contributing to the effective action. Although in making numerical estimates we are hampered by large number of unknown parameters in the 2HDM, it is worth remembering the existence of these effects as potential sources of enhancement in electroweak baryogenesis. \section{Acknowledgment} This work has been supported in part by the Department of Science and Technology, Government of India.
\section{Introduction} \label{intro} There are various ideas of generalization of the manifold notion. For example Aronszajn and Marshall \cite{1af,27af} have developed theory of so-called subcartesian spaces. Mostov \cite{28af} and Spallek \cite{36af,59af} have defined spaces called at present Mostov spaces. Roman Sikorski defined differential spaces (d-spaces for short) \cite{3bo}. There is a strong requirement for notions more general than manifolds in theoretical physics both classical and quantum . Here are few examples. \begin{itemize} \raggedright \item[ a)] Theory of singularities in General Relativity; space-time with its singular boundary is not a manifold \cite{33af,1bo}, \item[ b)] The space of solutions of Einstein equation (superspace) is not a manifold \cite{12bo}, \item[ c)] A class of non-differential processes described in book by Arnold \cite{13bo}. \end{itemize} The generalizations of the manifold concept have not found applications in physics yet. Among reasons of such situation I can mention excessive abstractness of some theories and small amount of theorems useful in practice. Differential spaces in the sense of Sikorski are generalization of the manifold concept by omitting the condition of existence of local diffeomorphisms to ${{\rm I\hspace{-0.5mm}R}}^n$. The generalization admit to considerations larger class of object than manifolds. Simultaneously, mathematical tools (definitions and theorems) of the d-spaces theory is similar to that one usually used in ordinary differential geometry. From the point of view of physics, where differential geometry plays a very important role, differential geometry on d-spaces is a natural tool for generalization of a physical theory. Other important argument which distinguishes the d-spaces theory is a great number of useful theorems. Among various structures existing in a space-time (e.g. topological, chronological, metrical, spinorial etc. ) only differential structure is not well elaborated. Since the structure may play a fundamental role in quantization of gravity \cite{14bo} every result concerning d-structure of space-time models seems to be valuable. \pagebreak \section{Differential spaces in the sense of Sikorski} \label{ds} A differential space in the sense of Sikorski is defined as follows. Let $(M,{\rm top}(M))$ be a topological space and \mbox{${\cal C}_0$} \ a set of real functions on $M$. A function \fun{f}{M}{{\rm I\hspace{-0.5mm}R}} is local \mbox{${\cal C}_0$} -function on $M$ if for every $p_0\in M$ there is a neighbourhood $U\in{\rm top}(M)$ and $\varphi\in\mbox{${\cal C}_0$}$ such that $f|_U=\varphi|_U$. The set of all local \mbox{${\cal C}_0$} \ functions on $M$ is denoted by $(\mbox{${\cal C}_0$})_{_M}$. \begin{defin} A set of real functions ${\cal C}$ on $M$ is said to be closed with respect to localization if ${\cal C}=({\cal C})_{_M}$. \end{defin} \begin{defin} A pair \dspa{M}{C} \ is said to be a differential space in the sense of Sikorski if: \begin{itemize} \label{defds} \raggedright \item[1)] ${\cal C}$ is closed with respect to localization, i.e. ${\cal C}=({\cal C})_{_M}$, \item[2)] ${\cal C}$ is closed with respect to superposition with smooth functions on ${{\rm I\hspace{-0.5mm}R}}^n$, $n\in{\rm I\hspace{-0.4mm}l\hspace{-1.1mm}N}$, i.e. for any $\ckow{\alpha}{1}{2}{n}\in {\cal C}$ and any $\omega\in C^{\infty}({\rm I\hspace{-0.5mm}R}^n)$, $\omega\circ (\ckow{\alpha}{1}{2}{n})\in {\cal C}$. \item[3)] $M$ is equipped with the topology ${\tau}_{_{\cal C}} $ which is the weakest topology on $M$ in which functions of ${\cal C}$ are continuous. \end{itemize} \end{defin} A picture of our world is formed by measures providing us with information coded in form of real functions. One can say that reality is given to us by a family of real functions existing on $M$. If the family is sufficiently "complete" one can identify it with the d-structure ${\cal C}$. Then the first axiom of definition \ref{defds} guarantees consistency of a local physics with the global one. The second axiom means that information obtained by superposition with smooth functions on ${{\rm I\hspace{-0.5mm}R}}^n$ do not lead to a new information not contained in ${\cal C}$ (see \cite{13af}). Let us test how the above described idea works in practice on example of cosmic string's space-time with a scalar field. \pagebreak \section{D-space of cosmic string's space-time} \label{cosmic} The space-time of cosmic string $(M,g)$ is a pseudoriemannian manifold with conical type quasiregular singularity and equipped with the metric \begin{equation} g=-dt^2+k^{-2}d\rho^2+{\rho}^2 d\phi^2 +dz^2 , \label{metric} \end{equation} where $k=(1-\Delta/2\pi)$ , $t, z \in {\rm I\hspace{-0.5mm}R}$, $\rho\in (0, \infty)$, $\phi\in \langle 0, 2\pi ) $ and $\Delta $ stands for the deficit of angle. This manifold is isometric to $(C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2, \iota^*\eta^{(5)})$, where $C^\circ$ is two-dimensional cone without vertex, $\iota : C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2$ $ \rightarrow{\rm I\hspace{-0.5mm}R}^5$ is an embedding and ${\eta}^{(5)}$ is the five-dimensional Minkowski metric. The singular boundary in this approach is represented by the set ${\rm S}\times {\rm I\hspace{-0.5mm}R}^2$, where ${\rm S}$ denotes vertex of the cone $C^{\bullet}$ and the dot under $C$ denotes that now the vertex is taken into account. The presented dipheomorphic picture of a space-time of cosmic string is interesting for two reasons. First, it shows the sense of the conical singularity in a demonstrative manner \cite{16af,15af}. Second, it enables us to construct immediately the differential structure in the sense of Sikorski for cosmic string's space-time both with and without singularity. In this approach, d-space of a cosmic string constitutes d-subspace $(C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2,$ $ ({\cal E}_5)_{C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2})$ of the d-space $({\rm I\hspace{-0.5mm}R}^5, {\cal E}_5 )$, where ${\cal E}_5 =C^{\infty} ({\rm I\hspace{-0.5mm}R}^5)$ and the symbol $(\cdot)_{C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2}$ denotes an operation of taking closure with respect to localization (definition \cite{3bo,12af,62af}). The cosmic string's space-time with singularity is not a manifold but it is still a d-space in the sense of Sikorski. The set $C^{\bullet}\times {\rm I\hspace{-0.5mm}R}^2 \subset {\rm I\hspace{-0.5mm}R} ^5$ can be treated as a d-subspace of the d-space $({\rm I\hspace{-0.5mm}R}^5, {\cal E}_5 )$. Then the d-space $(C^{\bullet}\times {\rm I\hspace{-0.5mm}R}^2, ({\cal E}_5)_{C^{\bullet}\times {\rm I\hspace{-0.5mm}R}^2})$ represents background d-space of cosmic string's space-time with singularity. The above mentioned d-spaces of cosmic string are not convenient from technical and methodological point of view. Therefore, in the following, two auxiliary d-spaces \tildedsp{P^\circ} \ and \tildedsp{P^\bullet} \ will be investigated rather than $(C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2, ({\cal E}_5)_{C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2})$ and $(C^{\bullet}\times {\rm I\hspace{-0.5mm}R}^2, ({\cal E}_5)_{C^{\bullet}\times {\rm I\hspace{-0.5mm}R}^2})$, respectively. The auxiliary d-spaces are defined as follows. \begin{defin} Let the set $\tilde{P}^\circ:={\rm I\hspace{-0.5mm}R}\times (0,\infty)\times \langle 0,2\pi\rangle \times {\rm I\hspace{-0.5mm}R}$ be a "parameter space". One can define the following d-structure $\mbox{$ \tilde{\cal P}^\circ$} :={\rm Gen}(\ckow{\tilde{{\alpha}}^{\circ}}{0}{1}{4})$, where functions \fun{{\tilde\alpha}^{\circ}_{i}}{\tilde{P}^\circ}{\rm I\hspace{-0.5mm}R}, $i=0,1,\dots ,4$ are given by means of the following formulas \begin{quote} \raggedright ${\tilde\alpha}^{\circ}_{0}(\tilde p):=t$ \linebreak ${\tilde\alpha}^{\circ}_{1}(\tilde p):=\rho\cos\phi$ \linebreak ${\tilde\alpha}^{\circ}_{2}(\tilde p):=\rho\sin\phi$\linebreak ${\tilde\alpha}^{\circ}_{3}(\tilde p):=z$\linebreak ${\tilde\alpha}^{\circ}_{4}(\tilde p):=\rho$, $\tilde{p}\in \tilde{P}^\circ $. \end{quote} \end{defin} The pair \tildedsp{P^\circ} \ is a d-space useful for description of interior of cosmic string's space time. \begin{defin} Let $\tilde{P}^\bullet:={\rm I\hspace{-0.5mm}R}\times \langle 0,\infty)\times \langle 0,2\pi\rangle \times {\rm I\hspace{-0.5mm}R}$, be the "prolonged parameter space". One can define the following d-structure $\mbox{$ \tilde{\cal P}^\bullet$} :={\rm Gen}(\ckow{{\tilde{\alpha}}^{\bullet}}{0}{1}{4})$. The functions \fun{\tilde{\alpha}^{\bullet}_{i}}{\tilde{P}^\bullet}{{\rm I\hspace{-0.5mm}R}} are defined as follows $$\tilde{\alpha}^{\bullet}_{i}(\tilde{p}_\bullet):= \lim_{\tilde{p}\rightarrow\tilde{p}_\bullet} {\tilde\alpha}^{\circ}_{i}(\tilde{p}), $$ where $\tilde{p}\in\tilde{P}^\circ,\tilde{p}_\bullet\in \tilde{P}^\bullet$ and $i= 0,1,\dots ,4$. \end{defin} The \tildedsp{P^\bullet} \ is used for description of a cosmic string's space time with singularity. Both \tildedsp{P^\circ} \ and \tildedsp{P^\bullet} \ are not Hausdorff topological spaces. Therefore they are not dipheomorphic to $(C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2, ({\cal E}_5)_{C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2})$ and $(C^{\bullet}\times {\rm I\hspace{-0.5mm}R}^2, ({\cal E}_5)_{C^{\bullet}\times {\rm I\hspace{-0.5mm}R}^2})$, respectively. However, after some identifications of points of \mbox{$ \tilde{P}^\circ$} \ and of \mbox{$ \tilde{P}^\bullet$} \ one can obtain d-spaces dipheomorphic to above mentioned background d-spaces of cosmic string's space-time (see \cite{16af,15af,62af} for details). In addition, every statement concerning smoothness and differential dimension true for \tildedsp{P^\circ} \ and \tildedsp{P^\bullet} \ holds also for $(C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2, ({\cal E}_5)_{C^{\circ}\times {\rm I\hspace{-0.5mm}R}^2})$ and $(C^{\bullet}\times {\rm I\hspace{-0.5mm}R}^2, ({\cal E}_5)_{C^{\bullet}\times {\rm I\hspace{-0.5mm}R}^2})$ (see \cite{62af} for details). Thus, without loss of correctness one can confine further considerations to \tildedsp{P^\circ} \ and \tildedsp{P^\bullet} . \section{Physical fields and differential structure for cosmic string} \label{scfield} Let us consider normal modes of a Klein-Gordon field on the cosmic string's space-time background. The modes have the following form: $$\fun{\tilde{\Psi}^{\circ}_{\epsilon,l,\beta}}{\tilde{P}^{\circ}}{\mbox{\rm\hspace{1.2mm},$$ \begin{equation} \label{psito} \tilde{\Psi}^{\circ}_{\epsilon,l,\beta}(t,\rho,\phi,z)= N_{\epsilon,l,\beta}e^{-i\epsilon t}e^{i\beta z}e^{il\phi}\rho^{|l|/k}F(l,k; \rho), \end{equation} where $\epsilon,\beta\in {\rm I\hspace{-0.5mm}R}$, $l\in\calk$, $N_{\epsilon,l,\beta}$ is the normalization constant and $k\in (0, 1) $ is defined in formula (\ref{metric}). The $F$ is an analytical function. Its detailed form is without meaning for further studies. \begin{defin} Let \dspa{M}{C} \ be a d-space. A complex function \fun{f}{M}{\mbox{\rm\hspace{1.2mm} \ is said to be smooth complex function if \ ${\rm Re}f$ and ${\rm Im}f$ are smooth ones. \end{defin} It is easy to check that \begin{stw} \label{smootho} For every $\epsilon, \beta\in{\rm I\hspace{-0.5mm}R}$, $l\in\calk$ and $k\in (0, 1)$ the normal modes \mbox{$\tilde{\Psi}^{\circ}_{\epsilon,l,\beta}$} \ are smooth functions on \tildedsp{P^\circ}. \end{stw} \begin{stw} \label{smoothb} For every $\epsilon ,\beta \in {\rm I\hspace{-0.5mm}R}$ and $l\in \calk$ the modes naturally prolonged to singularity \begin{equation} \label{psitb} \mbox{$\tilde{\Psi}^{\bullet}_{\epsilon,l,\beta}$} (p):=\lim_{q\rightarrow p}\mbox{$\tilde{\Psi}^{\circ}_{\epsilon,l,\beta}$} (q), q\in\mbox{$ \tilde{P}^\circ$}, p\in\mbox{$ \tilde{P}^\bullet$}, \end{equation} are \begin{itemize} \raggedright \item[{\bf a)}] smooth functions on \tildedsp{P^\bullet} \ for $k=1/n$, $n=1,2,\dots$, \item[{\bf b)}] not smooth functions on \tildedsp{P^\bullet} \ for other $k\in (0,1)$. \end{itemize} \end{stw} Following proposition \ref{smootho} one can say that the d-space of a cosmic string's gravitational field is from the very beginning "prepared" for insertion of a Klein-Gordon field on its background. The fact of insertion of the Klein-Gordon field is from the point of view of cosmic string's space time interior the only an indication which functions among already existing in \mbox{$ \tilde{\cal P}^\circ$} \ are the normal modes. The situation is much more interesting when one takes into account the conical singularity (proposition \ref{smoothb}). Then, in general, the insertion of the scalar field enforces a supplementation of \mbox{$ \tilde{\cal P}^\bullet$} . The situation clarifies the following proposition. \begin{stw} Let $\hat{\cal P}^{\bullet}$ be the smallest d-structure containing \mbox{$\tilde{\Psi}^{\bullet}_{\epsilon,l,\beta}$} \ as smooth functions; $\hat{\cal P}^{\bullet}:= {\rm Gen}(\ckow{\tilde{\alpha}^\bullet}{0}{1}{4},\mbox{$\tilde{\Psi}^{\bullet}_{\epsilon,l,\beta}$} )$. Then the d-space $(\mbox{$ \tilde{P}^\bullet$} ,\mbox{$ \hat{\cal P}^{\bullet}$} )$ has the following properties \begin{itemize} \item[{\bf a)}] $(\mbox{$ \tilde{P}^\bullet$} ,\mbox{$ \hat{\cal P}^{\bullet}$} ) = \tildedsp{P^\bullet}$ for $k=1/n$, $n=1,2,\dots$, \item[{\bf b)}] $(\mbox{$ \tilde{P}^\bullet$} ,\mbox{$ \hat{\cal P}^{\bullet}$} )$ \ is not dipheomorphic to \tildedsp{P^\bullet} \ for $k\ne 1/n$, $n=1,2,\dots$, \item[{\bf c)}] {\rm dim} $T_{p}(\mbox{$ \tilde{P}^\bullet$},\mbox{$ \hat{\cal P}^{\bullet}$}) = 6$ and {\rm dim} $T_{p}\tildedsp{P^\bullet} = 5$ for $p\in {\bf S}$, where {\bf S} denotes the set of singular points. \end{itemize} \end{stw} Details of proofs one can find in \cite{62af} . For $\Delta\ne 2\pi (1-1/n)$ the insertion of a Klein-Gordon field changes global properties of cosmic string's background d-space. The new background d-space $(\mbox{$ \tilde{P}^\bullet$},\mbox{$ \hat{\cal P}^{\bullet}$})$ is not dipheomorphic to \tildedsp{P^\bullet} . For example the dimension of its tangent spaces at singular points is 6 while for \tildedsp{P^\bullet} \ is 5. That means that $(\mbox{$ \tilde{P}^\bullet$},\mbox{$ \hat{\cal P}^{\bullet}$})$ do not represent background d-space of a cosmic string. In the case of $\Delta = 2\pi (1-1/n)$ the supplemented d-structure \mbox{$ \hat{\cal P}^{\bullet}$} \ and \mbox{$ \tilde{\cal P}^\bullet$} \ are the same so the background d-space of a cosmic string's space-time does not change. The similar result holds for an electromagnetic field and gravitational radiation on cosmic string's space-time background \cite{62af}. Thus, one can formulate the following theorem. \begin{tw} \label{nostringA} The conical space-time of cosmic string equipped with the metric {\rm (}\ref{metric}{\rm )} can be a background for a smooth Klein-Gordon field, electromagnetic field and a smooth field of gravitational radiation the only in the case of the following discrete spectrum of the deficit of angle: $$\Delta=2\pi (1-1/n), $$ where $n=1,2,\dots $\ . \end{tw} \section{Comparison with results obtained by means of field theory methods} \label{gradiation} The issues of previous section were obtained by means of strictly geometrical methods within the theory of differential spaces in the sense of Sikorski. There appear the question: whether an echo of these anticipations one can find among results obtained by the field theory methods? In the present section I will discuss some results of papers by Hacyan and Sarmiento \cite{44af} and also by Aliev and Gal'tsov \cite{43af,56af}. {\bf 1) } Sarmiento and Hacyan have studied the energy density spectrum of vacuum surrounding cosmic string for massless fields with spin 0, 1/2, and 1. Briefly speaking they have shown that the energy density of vacuum is given by the following formula \begin{equation} \label{vacuum} \frac{de}{d\omega}=\frac{\hbar {\omega}^{3}}{2{\pi}^{2}c^{3}}\left(1- \frac{1}{2\pi\omega k\rho}\sin (\frac{\pi}{k}){\rm I}_{_{\frac{1}{k}}}(2\omega\rho)\right), \end{equation} where $k$ has meaning as in (\ref{metric}). When $k=1/n$, $n=1,2,\ldots$ and so for deficits of angle $\Delta=2\pi (1-1/n)$ the energy density is like for the Minkowski space-time: $$\frac{de}{d\omega}=\frac{\hbar {\omega}^{3}}{2{\pi}^{2}c^{3}} . $$ One can say that in this case the energy density is not distorted. The result is independent of spin. {\bf 2) } The conical nature of space-time cosmic string geometry is a cause of a class of effects absent in the Minkowski space-time (for example: the lensing effect \cite{51af}). One of the most interesting effect is a gravitational radiation emitted by a freely moving particle (conical bremsstrahlung). The total energy emitted by the particle in the form of gravitational radiation is given by the formula \begin{equation} \label{radiation} E=\frac{1}{k}{\sin}^{2}(\frac{\pi}{k})f(v, k, d), \end{equation} where $f$ \ is a function of parameters of motion. Its explicit form one can find in the article by Aliev \cite{43af}. It is easy to check that the conical bremsstrahlung effect vanishes when $k=1/n$, $n=1,2,\ldots $ like in the above section. \section{Final remarks} Results concerning cosmic string's space-time obtained in previous sections could be interpreted as follows. The assumption of smoothness of physical fields introduces a new kind of coupling. Usually the coupling of a gravitational field and physical fields holds throughout the energy-momentum tensor in Einstein equations. The new coupling has a global nature and carries out through differential structure of cosmic string leading to changes in global properties of this space-time. When one enforce vanishing of such "differential coupling" then the gravitational field of a cosmic string becomes quantized and simultaneously every effect mentioned in the section \ref{gradiation} vanishes.
\section{Introduction} Multilayer films formed from transition metals and semiconductors have long been studied because of their unusual superconducting properties\cite{ruggiero} and because of possible application as x-ray optical elements.\cite{barbee85} Many unusual phenomena have been produced, ranging from the observation of dimensional crossover in weakly coupled superconducting Nb layers in Nb/Ge multilayers\cite{ruggiero} to the occurrence of bcc Ge in short-period Mo/Ge multilayers.\cite{wilson} Unusual magnetic properties have recently been observed in Fe/Si multilayers by workers at ETH\cite{toscano} and Argonne.\cite{fullerton} A large antiferromagnetic (AF) interlayer coupling in these multilayers manifests itself in hysteresis loops as a high saturation field and a low remanent magnetization. Similar magnetization curves are associated with large interlayer coupling in metal/metal multilayers like Fe/Cr and Co/Cu.\cite{demokritov,mosca} Much consideration has been given to whether the coupling in the Fe/Si system has the same origin as in the metal/metal multilayers.\cite{bruno,singh} Therefore the question of whether the spacer layer in the Fe/Si multilayers is a metal or semiconductor is of particular interest. Previous work on Nb/Si,\cite{fullerton2} Co/Si,\cite{miura} Ni/Si,\cite{wang} and Mo/Si\cite{stearns} multilayers have shown that there is a strong tendency towards compound formation at the metal/silicon interface. In general these multilayers consist of polycrystalline metal layers separated by an amorphous silicon layer which is bounded on either side by a layer of intermixed material. The intermixed silicide layers in these films were amorphous unless they were annealed at several hundred $^{\circ}$C.\cite{miura,holloway} These previously studied multilayers were therefore likely in their as-grown state to have metal/semiconductor character because of the presence of the amorphous silicon layer. In order to investigate the character of the spacer layer in the Fe/Si multilayer system, we have grown a large number of films with different substrate temperatures, substrate types, and layer thicknesses. When the Si spacer layer thickness is greater than about 20\AA, we find that the metal layers are crystalline but that the spacer layers are amorphous, similar to the situation in other transition metal/silicon systems. When the Si spacer layer thickness is less than about 20\AA thick, the iron silicide spacer layer forms a crystalline silicide with either the B2 or DO$_3$ structure. The B2 structure consists of two interpenetrating simple cubic sublattices and is identical to the CsCl structure for a 1:1 ratio of Fe and Si,\cite{vonkanel} while the DO$_3$ structure is an fcc lattice with two inequivalent Fe sites.\cite{kudrnovsky} Extensive growth experiments, described below, suggest that crystallinity of the spacer layer is crucial for occurrence of the antiferromagnetic interlayer coupling, in keeping with previous suggestions.\cite{fullerton} Since both the B2 and DO$_3$ phases are metallic,\cite{vonkanel,kudrnovsky} the fact that crystallinity is required for antiferromagnetic coupling suggests that the coupling in Fe/Si has a common origin with that observed in metal/metal multilayers. \section{Experimental Methods} The Fe/Si multilayers are grown in the ion-beam sputtering (IBS) chamber whose layout is shown schematically in Figure~\ref{chamber}. The system base pressure is typically about 2$\times$10$^{-8}$ torr. The ion gun is a 3 cm Kauffman source with focusing optics.\cite{commonwealth} The energy of the ions leaving the gun can be modulated by raising and lowering the voltage on the acceleration grids, creating in effect an electrical shutter. The Ar ions are incident on the sputter target at 1000V at an angle of about 45$^{\circ}$. The Ar pressure is maintained in the 2-3$\times$10$^{-4}$ range by a flow-controller coupled to a capacitance manometer.\cite{MKS} Four 3" diameter sputter targets are mounted on a tray which can be rotated by a stepper motor.\cite{MDC} Layer thickness is monitored by a quartz-crystal oscillator which is placed in close proximity to the substrates. The substrates are about 25 cm above the targets, clamped to a copper tray. The temperature of the tray is monitored by a thermocouple and can be varied between -150$^{\circ}$C and +200$^{\circ}$C.\cite{rsipaper} Three films are grown per chamber pumpdown. The thickness monitor, the controller for the stepper motor and the ion-beam power supply are all interfaced to a personal computer which has been programmed using the ASYST instrument control package.\cite{keithley} When the system is depositing a multilayer, the computer sends the material parameters to the thickness monitor, rotates the stepper motor to its new orientation, and turns the ion gun on. When the desired thickness is reached, the thickness monitor turns the ion-gun off and prompts the computer for the next layer. The basic design of the system is similar to one previously described by Kingon {\it et al.}\cite{kingon} The substrates for multilayers growth include glass coverslips, oxidized silicon wafers, MgO (001) and Al$_2$O$_3$(0\={2}11). The first two substrates, which are used for growth of polycrystalline films, were rinsed in solvents before loading into the vacuum chamber. The second two, which are used for epitaxial growth, are cleaned according to a recipe reported by Farrow and coworkers.\cite{farrow} The typical deposition rate for Fe is 0.2\AA/s while that for Si is about 0.3\AA/s. All films are capped with a 200\AA\ Ge oxidation barrier. The magnetic and structural properties of the films are stable for at least one year. Ge is used for capping instead of Si in order to prevent interference with element-specific soft x-ray fluorescence measurements, which will be reported elsewhere.\cite{carlisle} X-ray diffraction characterization has been performed using a 18kW rotating anode system outfitted with a graphite monochromator. All spectra are taken using the Cu K$_{\alpha}$ wavelength. Conventional high-resolution electron microscopy and electron diffraction have been performed in order to characterize the microstructure of the as-deposited films in cross-section. Magnetization curves are obtained using a vibrating sample magnetometer. All the data shown here were taken at room temperature. \section{Results} Overall the magnetic properties of the Fe/Si multilayers made by IBS are similar to those made previously by magnetron sputtering.\cite{fullerton,foiles} Definitive confirmation of AF interlayer coupling in our multilayers has been obtained by polarized neutron reflectivity measurements.\cite{ankner} For some unknown reason the magnetic properties of our multilayers are closer to those of the magnetron-sputtered multilayers than those reported on in a previous study using IBS, where much lower saturation fields were observed.\cite{inomata} The differences between the previous IBS-grown films and ours may be related to the lower ion-beam voltage used by Inomata {et al.}\cite{inomata} Comparisons on the basis of layer thickness are made here only between films grown during the same deposition run in order to insure that the relative layer thicknesses are meaningful. Films with similar layer thicknesses have been grown many times to establish reproducibility of the observed trends. \subsection{Layer-Thickness Dependence of Properties} \label{thicksec} Forty- and fifty-repeat multilayers have been grown with t$\rm_{Fe}$ = 14, 20, 30, 40, and 50\AA\ and t$\rm _{Si}$ = 14 and 20\AA. Magnetization curves for 50-repeat (Fe30\AA/Si20\AA) and (Fe30\AA/Si14\AA) multilayers grown on glass at nominal RT (about +60$^{\circ}$C) are shown in Figure~\ref{sithick}. On the y-axis of this plot is the magnetic moment of the multilayer normalized to the moment of an equivalent volume of bulk Fe. The magnetization curve of the (30/20) multilayer looks much like that of an Fe film, while the magnetization curve of the (30/14) multilayer shows the high saturation field and low remanence which characterize AF interlayer coupling. At its saturation field the magnetization of the (30/14) multilayer is about the same as for the (30/20) multilayer. Both of these films have moments only about half as large as an equivalent volume of bulk Fe. Our observation of AF coupling for Si thicknesses between 10 and 20\AA\ and the disappearance of coupling for Si thicker than 20\AA\ confirm previous observations on magnetron-sputtered films.\cite{fullerton} X-ray diffraction spectra for these multilayers are shown in Figures~\ref{sithickxrdlow} and ~\ref{sithickxrdhi}. Figure~\ref{sithickxrdlow} shows the small-angle x-ray scattering (SAXS) data with peaks at angles \begin{equation} n^2 \lambda^2 \, = \, 4 \Lambda^2 sin^2 \theta \; + \; 2\delta \label{lowbragg} \end{equation} \noindent where $\lambda$ is the x-ray wavelength, $\rm \Lambda \: \equiv \: t_{Fe} \: + \: t_{Si}$ is the multilayer bilayer period and $\delta$ is the index of refraction for x-rays.\cite{fullerton3} This grazing incidence data gives information about the quality of the multilayer interfaces. Figure~\ref{sithickxrdlow} shows four low-angle peaks for both films, indicating a reasonably strong composition modulation along the growth direction. (The higher frequency oscillations between 1$^{\circ}$ and 4$^{\circ}$ are finite-thickness fringes from the Ge cap layer.) The most notable difference between the two spectra is that the multilayer peaks are broader for the AF-coupled t$\rm _{Si}$ = 14\AA\ film, indicating more fluctuations in bilayer period and probably more interface roughness. Using the spacing between peak positions to eliminate the unknown $\delta$ from Eqn.~\ref{lowbragg} gives values of the bilayer period $\Lambda$ for the two films. For the multilayer with nominal layering of (Fe30\AA/Si20\AA)x50, the derived value for $\Lambda$ is (41.82 $\pm$ 0.07)\AA, while for the (Fe30\AA/Si14\AA)x50 film $\Lambda$ = (38.10 $\pm$ 0.04)\AA. $\Lambda$ is 8.2\AA\ shorter than the nominal value for the t$\rm _{Si}$ = 20\AA\ film, and 5.9\AA\ shorter than nominal for the t$\rm _{Si}$ = 14\AA\ film. Although some of the discrepancy between the nominal and observed bilayer period may be due to calibration inaccuracies, most is undoubtedly due to intermixing of the Fe and Si layers, in keeping with observations in the other metal/Si multilayers.\cite{fullerton2,holloway} Throughout this paper we will continue for convenience to refer to the films in terms of their nominal layer thicknesses. Comparison of the magnetization data to the x-ray data can give some further insight into the question of intermixing. Because of the presumed interdiffusion of the Fe and Si layers, the magnetic moment of the Fe layers is also reduced from the nominal value. The missing magnetic moment can be expressed as an equivalent thickness of Fe. Figure~\ref{missmom} shows a plot of missing moment in units of \AA ngstroms of Fe versus missing bilayer period determined from multilayer peak positions in SAXS for films grown at room temperature (RT). The plot shows that while the diffusion-induced reduction in bilayer period varies between 1 and 8\AA, the missing Fe moment per bilayer (for both interfaces) is consistently between 10 and 12\AA. The one outlier in Figure~\ref{missmom} is for a film which had t$\rm _{Fe}$ = 20\AA, the thinnest Fe for which we have ever observed interlayer coupling. Other groups have previously observed a moment reduction of 12-14\AA\ per bilayer in polarized neutron reflectivity measurements on uncoupled Fe/Si multilayers with thick Si layers.\cite{dufour,ankner2} The disparity between the magnetic moment reduction and the bilayer period reduction numbers may at first appear to be puzzling. This disparity occurs because the moment and bilayer period are affected by different aspects of the structure. In calculating the moment reduction in \AA\ the assumption has been made that the Fe layer has the magnetization of bulk Fe. This is equivalent to assuming that there is no Si in the Fe layer, which is undoubtedly false. In calculating the missing bilayer period, the assumption has also been made that the spacer layer is pure Si, also clearly false. The fact that the missing magnetic moment is almost constant irrespective of the reduction in bilayer period suggests that the spacer layer is non-magnetic independent of Si thickness. The lack of variation of the missing moment is then explained by the diffusion of a constant number of iron atoms into the silicon layer, irrespective of its thickness. The wide variation of the measured bilayer period is most likely related to the varying orientation and crystallinity of the spacer layer, neither of which affects the magnetic moment if the spacer itself is non-magnetic. Figure~\ref{sithickxrdhi} shows the high-angle x-ray spectra where peak positions give information about the orientation and crystallinity of the films. The intense peak near 70$^{\circ}$ in this plot is due to the Si substrate. Included are data for an (Fe40\AA/Si14\AA)x40 antiferromagnetically coupled multilayer and for an (Fe30\AA/Si20\AA)x40 uncoupled multilayer, both grown on oxidized Si(001) at RT. The peaks for the (40/14) film are narrower than for the (30/20). The Scherrer formula gives 78\AA\ or about two bilayer periods for the coherence length of the (40/14) film and 34\AA\ or about one bilayer period for the coherence length of the (30/20) film. Coherence lengths in IBS-sputtered antiferromagnetically coupled films are often as long as 200\AA. Fullerton {\it et al.} have inferred that the spacer layer in thin-Si Fe/Si multilayers must be crystalline based on their observation of coherence lengths longer than a bilayer period.\cite{fullerton} In keeping with its superior crystallinity, the (40/14) multilayer has one superlattice satellite on the low-angle side of the Fe(002) peak. Typically only one satellite on the low-angle side of the Fe (011) or (002) x-ray peak is observed for polycrystalline multilayers grown on glass, in agreement with observations by Foiles {et al.}\cite{foiles2} The thin-Si multilayers which have AF coupling usually show a mixed [001] and [011] orientation when grown on glass substrates at RT. Occasionally t$\rm _{Si}$ = 14\AA\ films with a pure (011) orientation are obtained at RT. The variation in texture may be due to changes in film stress under slightly different deposition conditions. Stress induced during deposition has been postulated to explain the mixed Mo texture found in Mo/Ge multilayers.\cite{wilson} In contrast to the thin-Si case, the thicker-Si Fe/Si multilayers which do not show interlayer coupling always have a pure (011) texture. Since the (011) plane is close-packed for the bcc crystal structure, one would expect the (011) orientation to be energetically favored for the Fe in a multilayer with amorphous Si. Films grown at nominal RT on glass or oxidized Si substrates typically had rocking curves about 10$^{\circ}$ wide indicating a moderate amount of orientation. Transmission electron microscopy (TEM) has been used to further investigate the morphology of the films. TEM cross-sectional images of an (Fe30\AA/Si20\AA)x50 multilayer and an (Fe40\AA/Si14\AA)x50 multilayer grown during the same deposition run are shown in Figures~\ref{TEMlowres}a and \ref{TEMlowres}b, respectively. The most salient features of the (30/20) multilayer are the long lateral continuity of the layers and the smoothness of the interfaces. Since there is no interlayer coherence in the (30/20) film, the crystalline grains have a high aspect ratio. The (40/14) multilayer also has long, continuous layer planes but has rougher interfaces, consistent with the SAXS data. Transmission electron selected-area diffraction patterns for the (30/20) and (40/14) films are shown in parts c and d of Figure~\ref{TEMlowres}. The (30/20) films shows only a Fe(011) ring, consistent with the high-angle x-ray diffraction scans. The (40/14) film, on the other hand, displays spots corresponding to the (011) and (002) reflections seen using x-rays. The presence of spots rather than rings in the (40/14) image implies the presence of large, oriented crystallites in the film. Most interestingly, the (40/14) image includes a faint spot near what would be the Fe(001) position were the Fe(001) peak not forbidden by symmetry in the bcc crystal structure. The (001) peak is allowed in the B2 and DO$\rm _3$ crystal structures. The B2 structure is found in the equilibrium phase diagram only at 10-22\% Si range of composition,\cite{binary} but workers at ETH have grown this crystal structure throughout the range of composition on Si substrates using MBE.\cite{onda} The DO$_3$ phase found in the equilibrium phase diagram is Fe$_3$Si, which is ferromagnetic.\cite{binary} Clearly a ferromagnetic spacer phase is not consistent with the observation of antiferromagnetic interlayer coupling, although a non-stoichiometric DO$_3$-structure phase might have different magnetic order. The B2 and $\epsilon$ iron silicide phases have both been previously suggested as possible candidates for the spacer layer in AF-coupled Fe/Si multilayers.\cite{fullerton,foiles2,dekoster} The position of the (001) TEM spots is not consistent with the $d$-spacings of the $\epsilon$ phase. According to the powder-diffraction files for the B2 and DO$_3$ structures, only the (111) peak of the fcc-family DO$_3$ does not coincide with a B2 peak. The (111) peak would be expected to be very weak in the diffraction patterns formed from cross-sectional specimens of the film. The reason is that a small number of grains contributes to the cross-sectional image, and the probability of sampling a grain with its (111) planes in the observable direction is small because of the random in-plane orientation. Future work will include electron diffraction studies of a (40/14) specimen prepared in the plan-view geometry, where the number of grains which are sampled is considerably larger and the odds of observing the fcc (111) peak are improved. High-resolution TEM images of the (30/20) and (40/14) multilayers are displayed in Figure~\ref{TEMhires}. The (30/20) film is shown in \ref{TEMhires}a to have a crystalline Fe layer and amorphous spacer layer, similar to the morphology seen before in Mo/Si\cite{stearns,holloway} and Co/Si multilayers.\cite{miura} The (40/14) multilayer in Fig.~\ref{TEMhires}b on the other hand is made up entirely of crystalline layers. The coherence between the Fe and silicide spacer is clearly evidenced by the continuity of atomic layer planes from the Fe layer into the spacer. Some crystallites in the (40/14) film extend all the way from the substrate to the surface of the film. The small coherence lengths observed in x-ray diffraction data for the uncoupled thicker-Si films are explained by the presence of the amorphous layers. The lack of crystallinity in the spacer layer of t$\rm _{Si}$ = 20\AA\ films is presumably due to insufficient time for full interdiffusion and ordering in the thicker layers. A kinetic mechanism for the lack of crystallization is supported by experiments which show that intentional placement of Fe in the Si layer allows thicker spacer layers to crystallize.\cite{foiles,mattson} Another striking feature of the image in Figure~\ref{TEMhires}b is the periodic modulation that occurs in the silicide spacer layer. The modulation originates from scattering by inequivalent planes of atoms. Simulation of this image using a multiple-scattering computer calculation may be helpful in positively identifying the crystal structure of the spacer layer phase. Dark-field images of the (40/14) multilayer can help answer questions about the texture of the film as well. Figure~\ref{TEMdark}a shows the same bright-field image as in Figure~\ref{TEMlowres}b. Dark-field images were formed using (001), (002) and (011) spots from the diffraction pattern shown in Figure~\ref{TEMlowres}d. The resulting micrographs are shown in Figures~\ref{TEMdark}b, c, and d respectively. Panels a and b of this figure show the same region of the (40/14) multilayer. The brightness of the spacer layers in this dark-field image demonstrates that the (001) reflection does indeed come from the spacer layer and is not the forbidden (001) spot of bcc Fe. Figures~\ref{TEMdark}c and d also show the same region (although a different region than panels a and b). The bright areas in these two images are the complement of one another; where one is bright, the other is dark and vice versa. The dark-field images in panels c and d of Figure~\ref{TEMdark} demonstrate convincingly that the orientation of the film evolves from predominantly (011) to predominantly (002) as the thickness increases. The reason for the change in orientation with film thickness is not obvious; it may be related to the bilayer-period-number dependence discussed in Section~\ref{bilayersec}. The effect of varying the Fe thickness has also been studied. Magnetic properties for films with 20\AA\ $\le$ t$\rm _{Fe}$ $\le$ 50\AA\ are found to change only slightly in keeping with the expected inverse proportionality of the saturation field with t$\rm _{Fe}$.\cite{demokritov} SAXS peaks tend to broaden and even split with increasing Fe thickness, indicating increased disorder in the layering. The splitting of these peaks may indicate different bilayer periods in areas of the film with the (011) and (001) textures. When the Fe is made less than 20\AA\ thick, the Fe high-angle diffraction peaks disappear and so does the AF coupling. The disappearance of crystalline Fe peaks near t$\rm _{Fe}$ = 20\AA\ is consistent with previous results on evaporated Fe/Si multilayers.\cite{dufour} Thus poor crystallinity of the Fe layers appears to suppress the interlayer coupling even when the Si thickness is favorable. The lack of AF coupling in films with poorly crystalline Fe may be related to the lack of a template for the crystalline iron silicide spacer to grow on. \subsection{Dependence of Properties on Growth Temperature and Post-Growth Annealing} \label{tempsec} Depositing the multilayers at different substrate temperatures is an obvious way of influencing the composition and crystallinity of the spacer layer phase in the Fe/Si multilayers. Fullerton has suggested that the interlayer of Fe/Si multilayers is improved by high-temperature growth.\cite{fullerton5} We have grown films on glass substrates at various temperatures between -150 and +200$^{\circ}C$. The effect of substrate temperature on the interlayer coupling of the films is illustrated in Figure~\ref{magtemp}, where magnetization curves for three (Fe40\AA/Si14\AA)x40 multilayers grown at -150$^{\circ}$C, +60$^{\circ}$C (nominal RT) and +200$^{\circ}$C are shown. The data show that as the substrate temperature increases the saturation field increases indicating larger AF coupling. The saturation magnetization also decreases, suggesting a larger degree of interdiffusion in the films grown at higher temperatures. The suspicion that more interdiffusion occurs at higher substrate temperatures is confirmed by examination of the SAXS spectra for the three films, shown in Figure~\ref{lowxrdtemp}. The film grown at reduced temperature has 7 peaks while the film grown at nominal RT has 5 and the film grown at +200$^{\circ}$C has only 4. Quantitative modelling of low-angle x-ray data has shown that the suppression of higher-order peaks may be due to either interdiffusion or cumulative roughness.\cite{fullerton3,payne} Certainly larger cumulative roughness could also occur at higher growth temperatures, but one would expect very rough growth to suppress AF coupling due to an increased number of pinholes and larger magnetostatic interlayer coupling.\cite{altbir} Since higher growth temperatures seem to enhance rather than suppress the coupling, it seems more likely that high substrate temperatures are promoting interdiffusion rather than roughness. Studies of Mo/Si multilayers showed that a growth temperature of 150$^{\circ}$C gives maximum SAXS reflectivity, which the authors attribute to greater interface smoothness than for RT deposition.\cite{stock} Smaller bilayer periods in multilayers grown at higher temperatures support the claim of increased interdiffusion. Fitting Eqn.~\ref{lowbragg} to peak positions from Figure~\ref{lowxrdtemp} gives $\Lambda$ = 52.7, 49.3, and 43.8 \AA\ respectively for the -150$^{\circ}$, +60$^{\circ}$, and +200$^{\circ}$ multilayers versus the nominal value of 54\AA. Higher substrate temperatures may also promote ordering of the Fe and Si atoms in the crystalline spacer layer. In the fully ordered B2 phase, the Fe and Si atoms sit on different simple cubic sublattices. The sublattice order can occur irrespective of whether or not the Fe to Si ratio is 1:1. It is interesting to speculate whether the AF coupling is dependent on the degree of ordering in the spacer layer. An ordering-dependent coupling seems plausible in light of the Fermi-surface theories of coupling in metal/metal multilayers.\cite{bruno2,stiles} A well-ordered B2 or DO$\rm _3$ phase would have more well-defined Fermi surface features than a random solid solution. Unfortunately the (001) silicide peak has only been observed by TEM, making experimental attempts to address this issue difficult. Further studies with x-ray diffraction and soft x-ray fluorescence are underway. The crystallinity of the films also varies with growth temperature. Surprisingly, films grown at both low and high temperatures on glass substrates always have only the (011) texture, while films grown at nominal RT often have mixed (001) and (011) textures. The multilayers deposited on heated and cooled substrates do differ greatly in that those grown at low temperature have amorphous spacer layers, while those grown at high temperatures have long crystalline coherence lengths. The reasons for the strange temperature dependence of growth texture are not understood, although one presumes that they have to do with the kinetics of growth. It is not clear why the (001) texture should appear at all, although it has also been seen in Mo/Ge multilayers.\cite{wilson} An oscillatory dependence of film texture on spacer layer thickness and deposition conditions has been reported for NiFe/Cu multilayers grown by IBS.\cite{nakatani} The (001) texture has not been reported in polycrystalline magnetron-sputtered Fe/Si multilayers, and may be due to some peculiarity of IBS growth. A logical extension to the growth temperature studies is to try annealing the Fe/Si multilayers grown at lower substrate temperatures to see if their properties evolve towards those of the multilayers grown at higher temperatures. As far as the magnetic properties are concerned, the answer is ``no.'' Annealing the uncoupled RT-grown (Fe30\AA/Si20\AA)x40 and low-temperature-grown (Fe40\AA/Si14\AA)x40 multilayers at +200$^{\circ}$C for two hours had almost no effect on their magnetic properties beyond a slight magnetic moment reduction. A subsequent 300$^{\circ}$C anneal for two hours once more produced a moment reduction and a decrease in coercive field in the uncoupled multilayers. A very low coercive field for annealed Fe/Si films is not surprising given the well-known softness of Fe-Si alloys. A 300$^{\circ}$C anneal even eliminated the interlayer exchange coupling of a RT-grown (Fe40\AA/Si14\AA)x50 film used as a control. For this (40/14) multilayer, the 300$^{\circ}$C anneal caused the SAXS peaks to narrow and reduced their number from 5 to 4. At the same time the bilayer period decreased from 49.4\AA\ to 46.0\AA. High-angle x-rays spectra (not shown) indicated that the Fe lattice constant slightly decreased, which is consistent with increased diffusion of Si in the Fe layer.\cite{foiles2} These x-ray and magnetization results imply that annealing primarily promotes interdiffusion of the Fe and silicide layers. With sufficient interdiffusion the spacer layer may become ferromagnetic, which would explain the suppression of antiferromagnetic interlayer coupling. These Fe/Si multilayers show less thermal stability than Mo/Si multilayers with comparable layer thicknesses, which do not show changes in SAXS spectra until 400$^{\circ}$C.\cite{stock} There was no sign of the solid-state amorphization previously observed in Fe/Si multilayers with thicker layers.\cite{gupta} Whatever process occurs during annealing, it does not enhance the interlayer coupling the way that +200$^{\circ}$C growth does. This is hardly surprising given that annealing will tend to drive the multilayer towards its equilibrium state, presumably a mixture of different iron silicide phases. There is no reason to think that the crystalline Fe/Fe$_x$Si$_{1-x}$ multilayer should be an intermediate phase during the annealing. In the future the kinetics of Fe/Si multilayer growth at different substrate temperatures will be investigated further by employing an ion-assist gun to improve atomic surface mobility. \subsection{Dependence of Properties on Number of Bilayers} \label{bilayersec} One puzzling aspect of the interlayer exchange coupling in the Fe/Si system has been the dependence of its strength on the number of bilayers in the multilayer. This trend is illustrated in Figure~\ref{binumber}, where magnetization curves for (Fe40\AA/Si14\AA)xN multilayers with 2, 12 and 25 repeats are displayed. (The 2-repeat multilayer is just an Fe/Si/Fe trilayer.) Although the trilayer has magnetic properties like bulk Fe, the 25-repeat multilayer data has a magnetization curve similar to the 40-repeat multilayer data shown above. The magnetization curve for the 12-repeat multilayer falls in between that for the thicker and thinner films. Evidence for AF coupling which is stronger near the top of an Fe/Si multilayer than near the substrate has previously been described by Fullerton {\it et al.}\cite{fullerton4} Presumably the increase of coupling with bilayer-number is a manifestation of the same phenomenon. The interlayer coupling in Co/Cu multilayers also increases with the number of bilayer periods up to about 25 bilayers.\cite{rupp} One would not expect interlayer coupling that is quantum-mechanical in nature to be affected much by total film thickness. The unusual thickness dependence therefore raises the question of whether there is quantum-mechanical coupling at all, or whether some other mechanism might determine the shape of the magnetization curves. Disordered magnetic materials such as small amorphous Fe particles can have low remanence and high saturation fields without any layering at all. The magnetization curves of these Fe particles are in fact quite similar to those of the Fe/Si multilayers.\cite{grinstaff} This resemblance might lead to speculation that the topmost Fe layers in Fe/Si multilayers are discontinuous and that the magnetic properties are dominated by particle shape. However, the existence of half-order peaks in polarized neutron reflectometry measurements in the IBS-grown Fe/Si multilayers\cite{ankner} and the magnetron-sputtered multilayers\cite{fullerton4} gives unambiguous evidence that the magnetic properties are due to magnetic order rather than structural disorder. In addition, TEM pictures such as Figure~\ref{TEMlowres} show that the Fe layers are continuous in films with both high and low saturation fields. How then does the number of bilayer periods influence the AF coupling strength? It has been suggested that the difference between thin and thick multilayers grown at nominal RT is that the substrates of thick multilayers have time to rise to a higher temperature (about +60$^{\circ}$C for our system) during the longer growth.\cite{fullerton5} This idea seems reasonable in light of the larger coupling in samples grown on heated substrates as described above. In order to investigate this idea, a (Fe100\AA/Si14\AA/Fe100\AA) film was grown on glass at +200$^{\circ}$C. The magnetization curve for this film is shown in Figure~\ref{trilayers}. Also shown in this figure are data for a (Fe100\AA/Si14\AA/Fe100\AA) trilayer deposited at nominal RT and for a (Fe100\AA/Si14\AA/Fe100\AA) trilayer deposited at +200$^{\circ}$C, both grown on a 500\AA-thick a-Si buffer. The trilayer deposited directly on glass at elevated temperature has only slightly less remanence and higher saturation field than the trilayer grown at RT whose data are shown in Figure~\ref{binumber}. This result implies that it is not substrate temperature alone which causes bilayer-number effects. The magnetization curves of the trilayers grown on buffer layers, on the other hand, look much more like typical t$\rm _{Fe}$ = 40\AA\ 40-repeat multilayer results. An epitaxial (Fe100\AA/Si14\AA/Fe100\AA) trilayer grown directly on an MgO(001) at +200$^{\circ}$C substrate also has strong AF coupling (data not shown). Undoubtedly the strong AF coupling of the trilayer grown directly on the MgO is due the superior surface quality of the single-crystal substrate. The take-away lesson from all of these results is that substrate roughness is probably responsible for the reduced interlayer coupling in (Fe40\AA/Si14\AA) multilayers with a low number of bilayers. Conformal growth may propagate this roughness up from the substrate into the multilayer. Parkin {et al.} have found that the interlayer coupling in MBE-grown Co/Cu multilayers is very sensitive to the substrate and the buffer layer type, perhaps due to pinholes through the Cu layers.\cite{parkin} Presumably thin Fe layers grown directly on glass are so wavy that pinhole and magnetostatic coupling dominate the interlayer interactions for the first few bilayer periods. Recent calculations show that magnetostatic effects associated with propagating roughness can give interlayer ferromagnetic coupling of the same order of magnitude as the coupling derived from quantum-well effects.\cite{altbir} Ongoing polarized neutron reflectivity experiments may give more information on the variation of the coupling with position in the thicker multilayers.\cite{ankner} \subsection{Growth on Single-Crystal Substrates} That Fe films can be grown epitaxially on MgO and $\rm Al_2O_3$ substrates is well-known.\cite{metoki} One might therefore expect to be able to grow high-quality Fe/Si superlattices on these substrates. Figure~\ref{epihigh}a) shows high-angle x-ray diffraction spectra for a purely (001)-oriented (Fe40\AA/Si\AA)x60 multilayer grown on MgO(001). The spectrum in Figure~\ref{epihigh}b) is data for a highly (011)-oriented (Fe40\AA/Si14)x46 multilayer grown on $\rm Al_2O_3$. Both multilayers were deposited at +200$^{\circ}$C. Figure~\ref{epihigh}c) shows a $\phi$ scan for the MgO (110) and Fe (110) peaks for the film on the MgO substrate. These sets of peaks are offset from one another by 45$^{\circ}$ in $\phi$, confirming the well-known epitaxial relation Fe(001) $\|$ MgO(001) and Fe(110) $\|$ MgO(100).\cite{metoki} The $\phi$ scans for the $\rm Al_2O_3$ substrate show that this film is only weakly oriented in-plane. Mattson {et al.} have previously grown Fe/FeSi multilayers on $\rm Al_2O_3$, but they did not comment on the orientation of the multilayer.\cite{mattson2} Rocking curves widths for both films are about 1$^{\circ}$ wide, indicating a considerably smaller mosaic than for the multilayers grown on glass. SAXS data for the multilayers on single-crystal substrates are comparable to the data for films grown on glass. The films grown on MgO are the only purely (001)-textured Fe/Si multilayers produced by IBS so far. Dekoster {et al.} have grown epitaxial Fe/FeSi multilayers on MgO(001) by MBE, but they do not present any x-ray diffraction data or magnetization curves.\cite{dekoster} Magnetization curves of films grown on single-crystal substrates (not shown) are qualitatively similar to those grown on glass or oxidized Si substrates. The only differences are that the saturation fields are higher for the epitaxial samples and that magnetocrystalline anisotropy effects are observed. The magnetocrystalline anisotropy energies of epitaxial trilayers grown on MgO and Ge are similar to bulk Fe.\cite{michel} The shape of the high-angle peaks plus superlattice satellites are described by a theory due to Fullerton {\it et al.}\cite{fullerton3} Application of this theory to the Fe/Si multilayers is difficult because the silicide lattice constant, the thickness of the remaining pure Fe and the thickness of the silicide spacer can be estimated only roughly. A precise determination of the silicide lattice constant should make a quantitative analysis of these satellite features possible. \section{Discussion} Fe and Si appear to be the only known transition-metal/semiconductor combination in which the two elements interdiffuse to form a crystalline spacer layer with coherent interfaces. The reasons why this unusual morphology occurs in the Fe/Si system are unknown but likely involve a high rate of Fe diffusion into a-Si and a low heat of crystallization of the iron silicide compound. A detailed discussion of these issues is beyond the scope of this paper. Three different crystal structures have been proposed for the crystalline spacer layer of the Fe/Si multilayers. The $\epsilon$ phase can be eliminated on the basis of the electron diffraction patterns and TEM dark field images presented here. The B2 and DO$_3$ crystal structures are better lattice-matched to Fe than $\epsilon$-FeSi or $\alpha$- and $\beta$-FeSi$_2$. The lattice constant of the B2 phase was reported by M\"ader and coworkers to be 2.77\AA, only 3.1\% different from Fe.\cite{mader} The lattice constant of the $\epsilon$ phase is 4.46\AA,\cite{mader} which matches the Fe(110) plane only in the energetically unfavorable (210) direction.\cite{mattson} Recent conversion-electron M\"ossbauer data are interpreted in support of the B2 crystal structure, although the possibility of the DO$_3$ phase was not considered in that study.\cite{dekoster} It is plausible that the B2 or DO$_3$ structures form in rapid, far-from-equilibrium growth conditions because of their small unit cells. Since silicon deposited at low substrate temperatures is amorphous, the most likely scenario is the following. Silicon deposited on a crystalline Fe layer goes down amorphous and diffuses only slightly into the Fe. Subsequently deposited Fe atoms diffuse rapidly into the amorphous Si, analogous to what happens during the growth of Mo/Si multilayers.\cite{stearns,holloway} During the diffusion of Fe into Si, crystallization of the silicide occurs, possibly driven by the heat of mixing or by the kinetic energy of the incident Fe atoms. Growth of the crystalline phase may proceed upward from the lattice-matched Fe template, or downward from the atomically bombarded film surface. If the growth of the crystalline silicide phase proceeds downward from the film surface, one might expect to see some crystalline silicide in the high-resolution TEM image for the t$\rm _{Si}$ = 20\AA\ film (Figure~\ref{TEMhires}b). The lack of any evidence for crystalline silicide in this image suggest that the crystallization proceeds upward from the iron/silicide interface, not downward from the film surface. It is difficult to determine how realistic this model for growth of the crystalline silicide is since the Fe/Fe-Si and Si-Fe/Fe interfaces appear identical in Figure~\ref{TEMhires}b. In contrast, the Mo/Si and Si/Mo interfaces in Mo/Si multilayers appear quite different from one another.\cite{stearns,holloway} In the Mo/Si multilayers, an amorphous MoSi$_2$ region appears which is thicker at the Mo/Si interface than at the Si/Mo interface. Detailed TEM studies of multilayers with t$\rm _{Si}$ larger than 20\AA\ may help to answer whether amorphous silicides can occur in IBS-grown Fe/Si multilayers. Using the B2 phase lattice constant reported by the Z\"urich group,\cite{mader} we can estimate the expected bilayer period of a nominal Fe/Si multilayer in which Fe atoms diffuse into the Si layer up to a 1:1 stoichiometry. The spacing between the Fe and Fe$\rm _x$Si$\rm _{1-x}$ layers is taken as the average of the interplanar spacings of the two materials. The result of this rough calculation is that an (Fe40\AA/Si14\AA) multilayer which interdiffuses up to the 1:1 stoichiometry should form a (Fe33.2\AA/FeSi16.3\AA) multilayer with a bilayer period of 49.4\AA. The missing bilayer period predicted from this model is 4.6\AA, in the middle of values on the x-axis of Figure~\ref{missmom}. One can also calculate the expected magnetic moment reduction assuming that Fe atoms in the silicide layer have no moment and those in the Fe layer have their full moment. Under this assumption a calculation predicts 8.2\AA\ of missing Fe moment, slightly lower than indicated in Figure~\ref{missmom}. This calculation neglects the possibility that some Fe atoms in the Fe layer with Si near neighbors may have reduced magnetic moments. In the discussion above the possibility has not been mentioned that the missing bilayer period and magnetic moment are due to an inaccurate thickness calibration. This explanation is contradicted by magnetization and x-ray diffraction measurements on Fe/Ge multilayers, where measured magnetic moments and bilayer periods are in much closer agreement with nominal values than for Fe/Si.\cite{michel} The improved agreement in the case of Fe/Ge multilayers suggests that interdiffusion is less important in multilayers with Ge spacer layers than in multilayers with Si spacers. The main point is that the formation of the B2 silicide does qualitatively explain the bilayer period reduction observed in the Fe/Si multilayers. The underlying reason for the bilayer period reduction is that the silicide which forms is denser than both Fe and Si. This situation is similar to that observed in other metal/Si multilayers\cite{fullerton2,holloway} except that in the other multilayers the silicide remains amorphous. Confirmation that the spacer layer phase has the B2 or DO$\rm _3$ structure is important for understanding the coupling mechanism in these compounds. Both the B2 and DO$\rm _3$ phases are known to be metallic for some ranges of composition.\cite{vonkanel,kudrnovsky} Thus the present results and those of other workers\cite{fullerton,dekoster} suggest that Fe/Si is really a metal/metal multilayer. The origin of the interlayer coupling is then likely to be described by the same theories as describe coupling in Co/Cu and Fe/Cr multilayers.\cite{bruno2,stiles} Fe/Si multilayers may therefore not be a good test case for theories which model interlayer exchange coupling across insulators.\cite{bruno,singh} In the discussion above the possibility has been neglected that the amorphous spacer layer in the thick-Si films may also be metallic. If both the thick amorphous spacers and the thin crystalline spacers are metallic silicides, then it must be the crystallinity that is the essential feature for the existence of AF interlayer coupling. Up to now there have been no reports of AF coupling across amorphous metallic spacer layers. Toscano {\rm et al.} have reported AF coupling across amorphous silicon spacer layers.\cite{toscano} These Fe/a-Si/Fe trilayers were prepared at low temperature so as to suppress interdiffusion.\cite{toscano} The character of AF coupling in the a-Si spacer trilayers is likely quite different than in the multilayers described in this study, where substrate heating increases the strength of coupling. At the moment there is no direct evidence regarding the metallic or insulating nature of the amorphous spacer layers found in the (Fe30\AA/Si20\AA) multilayers. Temperature-dependent current-in-plane resistivity measurements suggest that both crystalline and amorphous spacer layers in Fe/Si multilayers are poorly conducting.\cite{michel} Fe$_{70}$Si$_{30}$ and Fe$_{65}$Si$_{35}$ amorphous alloys have a temperature-independent resistivity, suggesting non-metallic behavior.\cite{marchal} Overall the evidence suggests that the amorphous spacer layers in (Fe30\AA/Si20\AA) multilayers are not metallic, but spectroscopic measurements like soft x-ray fluorescence\cite{carlisle} are needed for confirmation. The interesting question as to whether there can be AF interlayer coupling across an amorphous metal spacer layer must then be left for another study. \section{Conclusions} An extensive study of the growth of Fe/Si multilayers by ion-beam sputtering has been performed. The crystalline quality of the films is better when they are grown with thick Fe layers, with thin Si layers, at high temperature, and on single-crystal substrates. Improved growth conditions lead to higher saturation fields and lower remanence in magnetization curves. Measured bilayer periods are consistently shorter in these multilayers than the nominal value, suggesting formation of a dense silicide phase in the spacer layer. Despite considerable interdiffusion in the multilayers, a strong composition modulation along the growth direction is maintained as evidenced by SAXS measurements. There are two surprising results from this study. One is that the films grow on glass with a mixed (011) and (001) texture near nominal RT and with a pure (011) texture at higher and lower temperatures. The other surprise is that the strength of the interlayer coupling depends strongly on the number of bilayer periods in films with thin Fe layers. This latter result is explained on the basis of substrate surface roughness. Unraveling the behavior of the Fe/Si multilayer system has proven to be a considerably more complex task than understanding the Fe/Cr or Co/Cu multilayer systems. The reason is that compound formation at the Fe/Si interface is crucial to understanding the AF interlayer coupling. Identification of possibly disordered phases in the spacer layer of a multilayer continues to be an experimental challenge. Mounting evidence suggests that the spacer layer in the AF-coupled Fe/Si multilayers is metallic and crystalline and that the Fe/Si interlayer coupling therefore has the same origin as in metal/metal multilayers. \bigskip \paragraph*{Acknowledgements} \noindent We would like to thank P.E.A. Turchi, T.W. Barbee Jr., T.P. Weihs, E.E. Fullerton, Y. Huai and E.C. Honea for helpful discussions, and B.H. O'Dell and S. Torres for technical assistance. Further thanks go to C.-T. Wang of Stanford for the four-circle x-ray diffractometry and to Sandia National Lab for use of their electron microscope for HREM work. Part of this work was performed under the auspices of the U.S. Department of Energy by LLNL under contract No. W-7405-ENG-48. \clearpage
\section{Introduction} Recently the CLEO collaboration has observed\cite{cleo1} the exclusive radiative decay $B \rightarrow K^* \gamma $ with a branching fraction of $BR(B\rightarrow K^* \gamma)=(4.5 \pm1.0 \pm0.9)\times10^{-5}$. The inclusive $b\to s \gamma$ branching ratio measured by CLEO\cite{cleo2} is: \begin{eqnarray} BR(B\to X_s \gamma ) = (2.32\pm 0.57 \pm 0.35) \times 10^{-4}. \end{eqnarray} The newest upper and lower limits of this decay branching ratio are \begin{eqnarray} 1.0\times 10^{-4} < BR(B \to X_s \gamma) < 4.2\times 10^{-4}, \ \ at \ \ 95\%C.L.. \end{eqnarray} As a loop-induced flavor changing neutral current(FCNC) process the inclusive decay(at quark level) $b\to s \gamma$ is in particular sensitive to contributions from those new physics beyond the Standard Model(SM)\cite{Hew1}. There is a vast interest in this decay. The decay $b\to s \gamma$ and its large leading log QCD corrections have been evaluated in the SM by several groups \cite{Ciu}. The reliability of the calculations of this decay is improving as partial calculations of the next-to-leading logarithmic QCD corrections to the effective Hamiltonian\cite{AJB,lcd}. On the other hand the discovery of the top quark and the measurement of its mass (in this paper we use the weighted average $m_t=180\pm 12\; GeV$ from the announced results of $m_t$ by CDF and D0\cite{cdfd0} wherever possible) at FERMILAB basically eliminated a source of uncertainties for the calculation of the decay $b\to s \gamma$ in the SM and in beyond theories. The great progress in theoretical studies and in experiments achieved recently encourage us to do more investigations about this decay in Technicolor theories. In this paper, we estimate the possible contributions to the decay $b\to s \gamma$ from the exchange of the charged Pseudo-Goldstone bosons which will appear in no-minimal Technicolor models, such as the Farhi-Susskind one-generation Technicolor model (OGTM) \cite{Farhi}. We know that the experimental data seems disfavor the OGTM which generally tend to predict $S$ parameter large and positive \cite{Peskin92}. Why we here still choose it to do the calculations? The reasons are the following: (1) At first, presence of the Pseudo-Goldstone bosons in the particle spectrum is a common feature of those non-minimal TC models with ordinary or novel ETC sectors, no matter the specific differences of structures between those models. The gauge couplings of the PGBs are determined by their quantum numbers, while the Yukawa couplings of PGBs to ordinary fermions are generally proportional to fermion masses for many TC/ETC models. Among the non-minimal TC models, the OGTM \cite{Farhi} is the simplest and most frequently studied model. Many relevant works \cite{King95} have been done since the late 1970's. One can use those existed results directly in further investigations. (2) On the other hand, the constraints on the $S$ parameter could be relaxed considerably by introducing three additional parameters $(V, W, X)$ \cite{Burgess94a}. A global fit to the data in which all six oblique parameters S through X are allowed to vary simultaneously gives the one standard deviation bound on $S$: $ S \sim -0.93 \pm 1.7$ \cite{Burgess94b}. This fact means that the constraint on the OGTM from the parameter $S$ could be considerably weakened if we consider the effects from light technifermions and light PGBs \cite{Evans94}. In this paper, we estimate the possible contributions to the rare decay $b\to s \gamma$ from the charged PGBs in the framework of the OGTM. At least, one can regard our results as an estimation for the ``correct'' output of the future ``realistic'' TC models. This paper is organized as the following: In Section 2, we present the basic ingredients of the OGTM and then calculate the PGB contributions to $b\to s \gamma $ decay, together with the full leading log QCD corrections. In Section 3, we obtain the branching ratios of this decay, and derive out the constraints on masses of charged PGBs by phenomenological analysis. The conclusions are also included in this section. \section{Charged PGBs and QCD Corrections to $b\to s \gamma$} In the OGTM \cite{Farhi}, when the technifermion condensate $<\overline{T}T>\neq 0$ was formed, the global flavor symmetry will break as follows: $SU(8)_L \times SU(8)_R \rightarrow SU(8)_{L+R}$. Consequently, 63 (Pseudo)-Goldstone bosons will be produced from this breaking. When all other interactions but the Technicolor are turned off, these 63 Goldstone bosons are exactly massless. Three of them are eaten by the $W^{\pm}$ and $Z^0$ gauge bosons. The others acquire masses when one turned on the gauge interactions, and therefore they are Pseudo-Goldstone Bosons(PGBs). According to previous studies, the phenomenology of those color-singlet charged PGBs in the OGTM is very similar with that of the elementary charged Higgs bosons $H^{\pm }$ of Type-I Two-Higgs-Doublet Model(2HDM) \cite{Gunion}. And consequently, the contributions to the decay $b\rightarrow s\gamma$ from the color-singlet charged PGBs in the OGTM will be very similar with that from charged Higgs bosons in the 2HDM. As for the color-octet charged PGBs, the situation is more complicated because of the involvement of the color interactions. Other neutral PGB's don't contribute to the rare decay $b\to s \gamma$. The gauge couplings of the PGBs are determined by their quantum numbers. The Yukawa couplings of PGBs to ordinary fermions are induced by ETC interactions and hence are model dependent. However, these Yukawa couplings are generally proportional to fermion masses with small differences in the magnitude of the coefficients for different TC/ETC models. The relevant couplings needed in our calculation are directly quoted from refs.\cite{Ellis,Chivukula95,Xiao94} and summarized in Table \ref{table1}, where the $V_{ud} $ is the corresponding element of Kobayashi-Maskawa matrix. For the OGTM, the Goldstone boson decay constant $F_\pi$ in Table \ref{table1} should be $F_{\pi}=v/2=123\;GeV$, in order to ensure the correct masses for the gauge bosons $Z^0$ and $W^{\pm}$ \cite{King95}. \begin{table}[htbp] \caption{ The relevant gauge couplings and Effective Yukawa couplings for the OGTM. }\label{table1} \begin{center} \vspace{.1cm} \begin{tabular}{|c|c|} \hline $P^+ P^- \gamma_\mu$ & $-ie(p_+ - p_-)_\mu$ \\ \hline $P^+_{8a} P^-_{8b} \gamma_\mu$ & $-ie(p_+ - p_-)_\mu \delta_{ab}$ \\ \hline $P^+\; u\; d$ & $i\frac{V_{ud}}{2 F_\pi}\sqrt{\frac{2}{3}} [M_u (1-\gamma_5) - M_d (1 + \gamma_5) ]$ \\ \hline $P^+_{8a}\; u\; d$ & $i\frac{V_{ud}}{2 F_\pi} 2 \lambda_a [M_u (1-\gamma_5) - M_d (1 + \gamma_5) ]$ \\ \hline $P^+_{8a} P^-_{8b} g_{c\mu}$ & $-g f_{abc}(p_a - p_b)_\mu $ \\ \hline \end{tabular} \end{center} \end{table} In ref.\cite{Randall}, Randall and Sundrum have estimated the contributions to $b\to s \gamma$ from the exchange of ETC gauge bosons in various ETC scenarios. In the case of ``traditional'' ETC (just the case which will be studied here), the dominant contribution to $b\to s \gamma$ occurs when the ETC gauge boson is exchanged between purely left-handed doublets and when the photon is emitted from the technifermion line. But the resulted ETC contribution is strongly suppressed with respect to the SM by a factor of $m_t/(4\pi v)< 0.09$ for $m_t< 200\;GeV$\cite{Randall}. In short, the ETC contribution to the decay $b\rightarrow s\gamma$ is small and will be masked by still large experimental and theoretical uncertainties. We therefore can neglect the ETC Contributions to $b\to s \gamma$ at present phenomenological analysis. In Fig.1, we draw the relevant Feynman diagrams which contribute to the decay $b\to s \gamma$, where the half-circle lines represent the W gauge boson of SM as well as the charged PGBs $P^{\pm}$ and $P_8^{\pm}$ of OGTM. In the evaluation we at first integrate out the top quark and the weak W bosons at $\mu=M_W$ scale, generating an effective five-quark theory. By using the renormalization group equation, we run the effective field theory down to b-quark scale to give the leading log QCD corrections, then at this scale, we calculate the rate of radiative $b$ decay. After applying the full QCD equations of motion\cite{eom}, a complete set of dimension-6 operators relevant for $b\to s \gamma $ decay can be chosen to be: \begin{eqnarray} O_1&=&(\overline{c}_{L\beta} \gamma^{\mu} b_{L\alpha}) (\overline{s}_{L\alpha} \gamma_{\mu} c_{L\beta})\;,\\ O_2&=&(\overline{c}_{L\alpha} \gamma^{\mu} b_{L\alpha}) (\overline{s}_{L\beta} \gamma_{\mu} c_{L\beta})\;,\\ O_3&=&(\overline{s}_{L\alpha} \gamma^{\mu} b_{L\alpha}) \sum_{q=u,d,s,c,b}(\overline{q}_{L\beta} \gamma_{\mu} q_{L\beta})\;,\\ O_4&=&(\overline{s}_{L\alpha} \gamma^{\mu} b_{L\beta}) \sum_{q=u,d,s,c,b}(\overline{q}_{L\beta} \gamma_{\mu} q_{L\alpha})\;,\\ O_5&=&(\overline{s}_{L\alpha} \gamma^{\mu} b_{L\alpha}) \sum_{q=u,d,s,c,b}(\overline{q}_{R\beta} \gamma_{\mu} q_{R\beta})\;,\\ O_6&=&(\overline{s}_{L\alpha} \gamma^{\mu} b_{L\beta}) \sum_{q=u,d,s,c,b}(\overline{q}_{R\beta} \gamma_{\mu} q_{R\alpha})\;,\\ O_7&=&(e/16\pi^2) m_b \overline{s}_L \sigma^{\mu\nu} b_{R} F_{\mu\nu}\;,\\ O_8&=&(g/16\pi^2) m_b \overline{s}_{L} \sigma^{\mu\nu} T^a b_{R} G_{\mu\nu}^a\;. \end{eqnarray} The effective Hamiltonian appears just below the W-scale is given as \begin{equation} {\cal H}_{eff} =\frac{4G_F}{\sqrt{2}} V_{tb}V_{ts}^* \displaystyle{\sum_{i=1}^{8} }C_i (M_W^-) O_i(M_W^-). \end{equation} The coefficients of 8 operators are calculated from diagrams of Fig.1: \begin{eqnarray} C_i(M_W)&=&0, \;\; i=1,3,4,5,6, \;\;\; C_2(M_W)=-1,\\ C_7(M_W)&=& A(\delta) +\frac{B(x)}{3\sqrt{2}G_F F_{\pi}^2 } +\frac{8 B(y)}{3\sqrt{2}G_F F_{\pi}^2 }, \label{c7}\\ C_8(M_W)&=& C(\delta) +\frac{D(x)}{3\sqrt{2}G_F F_{\pi}^2 } +\frac{8 D(y) + E(y)}{3\sqrt{2}G_F F_{\pi}^2 },\label{c8} \end{eqnarray} with $\delta=M_W^2/m_t^2$, $x=(m(P^{\pm})/m_t)^2$ and $y=(m(P^{\pm}_8)/m_t)^2$. The functions $A$ and $C$ arise from graphs with W boson exchange are already known contributions from SM; while the functions $B$, $D$, and $E$ arise from diagrams with color-singlet and color-octet charged PGBs of OGTM. They are given by, \begin{eqnarray} A(\delta)&=& \frac{ \frac{1}{3} +\frac{5}{24} \delta -\frac{7}{24} \delta^2}{(1-\delta)^3} +\frac{ \frac{3}{4}\delta -\frac{1}{2}\delta^2}{(1-\delta)^4} \log[\delta] \\ B(y)& =& \frac{ -\frac{11}{36} +\frac{53}{72}y -\frac{25}{72}y^2}{(1-y)^3} +\frac{ -\frac{1}{4}y +\frac{2}{3}y^2 -\frac{1}{3}y^3} {(1-y)^4}\log[y],\\ C(\delta)&=& \frac{\frac{1}{8} -\frac{5}{8} \delta-\frac{1}{4} \delta^2}{(1-\delta)^3} -\frac{ \frac{3}{4}\delta^2}{(1-\delta)^4} \log[\delta] \\ D(y)& =& \frac{ -\frac{5}{24} +\frac{19}{24}y -\frac{5}{6}y^2}{(1-y)^3} +\frac{ \frac{1}{4}y^2 -\frac{1}{2}y^3}{(1-y)^4} \log[y],\\ E(y) & =& \frac{ \frac{3}{2}-\frac{15}{8}y -\frac{5}{8}y^2 }{(1-y)^3} +\frac{\frac{9}{4}y -\frac{9}{2}y^2}{(1-y)^4 }\log[y]. \end{eqnarray} It is shown from these expressions that, for $\delta < 1$, $x,y >> 1$, the OGTM contribution $B$, $D$ and $E$ have always a relative minus sign with the SM contribution $A$ and $C$. As a result, the OGTM contribution always destructively interferes with the SM contribution. This can also be seen from the numerical results and discussion in the next section. The running of the coefficients of operators from $\mu=M_W$ to $\mu=m_b$ was well described in refs.\cite{Ciu}. After renormalization group running we have the QCD corrected coefficients of operators at $\mu=m_b$ scale. \begin{equation} C_7^{eff}(m_b) = \eta^{16/23}C_7(M_W) +\frac{8}{3} ( \eta^{14/23}-\eta^{16/23} ) C_8(M_W) +C_2(M_W) \displaystyle \sum _{i=1}^{8} h_i \eta^{a_i}. \end{equation} With $\eta = \alpha_s(M_W) /\alpha_s (m_b)$, $$ h_i =\left( \frac{626126}{272277}, -\frac{56281}{51730}, -\frac{3}{7}, -\frac{1}{14}, -0.6494, -0.0380, -0.0186, -0.0057 \right),$$ $$a_i = \left( \frac{14}{23}, \frac{16}{23}, \frac{6}{23}, -\frac{12}{23}, 0.4086, -0.4230, -0.8994, 0.1456 \right).$$ \section{The $B \rightarrow X_s \gamma$ decay rate and phenomenology} Following refs.\cite{Ciu}, applying a spectator model, \begin{eqnarray} BR(B \rightarrow X_s \gamma) /BR(B\rightarrow X_c e\overline{\nu}) \simeq\Gamma(b\rightarrow s\gamma)/\Gamma (b\rightarrow ce\overline{\nu}). \end{eqnarray} Then when have \begin{eqnarray} \frac{BR(B \rightarrow X_s \gamma)}{BR(B \rightarrow X_c e \overline{\nu})} \simeq \frac{|V_{tb} V_{ts}^*|^2} {|V_{cb}|^2} \frac{6 \alpha_{QED}}{\pi g (m_c/m_b)} |C_7^{eff}(m_b)|^2 \left(1-\frac{2 \alpha_{s}(m_b)}{3 \pi} f(m_c/m_b) \right)^{-1}, \end{eqnarray} where the phase space factor $g(z)$ is given by: \begin{eqnarray} g(z)=1-8z^2+8z^6-z^8-24z^4\log z, \end{eqnarray} and the factor $f(m_c/m_b)$ of one-loop QCD correction to the semileptonic decay is, \begin{eqnarray} f(m_c/m_b)=(\pi^2 -31/4)(1-m_c^2/m_b^2) + 3/2. \end{eqnarray} Afterwards one obtains the $B \rightarrow X_s \gamma$ decay rate normalized to the quite well established semileptonic decay rate $Br(B \to X_c e\overline{\nu} )$. If we take experimental result $BR(B \to X_c e\overline{\nu} ) =10.8\% $\cite{data}, the branching ratios of $B \to X_s \gamma$ is found to be: \begin{eqnarray} BR(B \rightarrow X_s \gamma) \simeq 10.8\%\times \frac{|V_{tb} V_{ts}^*|^2} {|V_{cb}|^2} \frac{6 \alpha_{QED}\;|C_7^{eff}(m_b)|^2} {\pi g (m_c/m_b)} \left(1-\frac{2 \alpha_{s}(m_b)}{3 \pi} f(m_c/m_b) \right)^{-1}. \end{eqnarray} In numerical calculations we always use $M_W=80.22\;GeV$, $\alpha_s(m_Z)=0.117$, $m_c=1.5\;GeV$, $m_b=4.8\;GeV$ and $|V_{tb} V_{ts}^*|^2/ |V_{cb}|^2= 0.95$ \cite{data} as input parameters. Generally speaking, the contribution to the decay $b\to s \gamma$ from color singlet $P^\pm$ is small when compared with the contribution from the color octet $P_8^\pm$, since there is a color enhancement factor $8$ appeared in the third terms in eqs.(\ref{c7}, \ref{c8}) for the functions $B(y)$ and $D(y)$. Fig.2 is the plot of the branching ratio $Br(B \rightarrow X_s \gamma)$ as a function of the top quark mass. The upper dashed curve in Fig.2 represents the branching ratio in the standard model, while the solid curve shows the same ratio with the inclusion of the contributions from $P^{\pm}$ and $P_8^{\pm}$ assuming $m(P^\pm) =300\;GeV$ and $m(P^\pm_8) =600\;GeV$. The band between two dash-dotted lines corresponds to the newest CLEO limits: $ 1.0\times 10^{-4} < BR(B \to X_s \gamma) < 4.2\times 10^{-4}$ at $95\%C.L.$ \cite{cleo2}. The branching ratio of $b \to s\gamma$ with large contribution from OGTM, is much more sensitive with the top quark mass, compared with the case of pure SM. It is known from the decoupling theorem that for heavy enough nonstandard boson, we should recover the SM result. So for sufficiently large values of $m(P^{\pm})$, $m(P^{\pm}_8)$ ( e.g. $m(P^{\pm})>600$GeV, $m(P^{\pm}_8)>2000$GeV), the contributions from OGTM shall be negligible. This can also be seen from the fact that the functions $B$, $D$ and $E$ go to zero, as $x$,$y\to \infty$. For not so large $m(P^{\pm})$, $m(P^{\pm}_8)$, the OGTM contribution cancels much of the SM contribution because of the relative minus sign between their contribution. As a result, the branching of $b\to s\gamma$ reached the lower limit of the CLEO experiment. So a large region of $m(P^{\pm})$, $m(P^{\pm}_8)$ ( i.e. $1000GeV <m(P^{\pm}_8)<2000GeV$, for all $m(P^{\pm})$ ) is ruled out. When $m(P^{\pm})$, $m(P^{\pm}_8)$ go on smaller, their contribution is about two times as large as contribution of SM (recall there is a relative minus sign), the branching ratio of $b \to s\gamma$ resumes to experiment allowed region. But if the $m(P^{\pm})$, $m(P^{\pm}_8)$ are smaller enough, the contribution of OGTM is more larger, the region is also excluded by the upper limit of CLEO experiment. The whole result is illustrated at Fig.3, large part of the parameter space in the $m(P^\pm)-m(P_8^\pm)$ plane can be excluded according to the current $CLEO$ 95\% C.L. limits on the ratio $BR(B \rightarrow X_s \gamma)$ \cite{cleo2}. It is easy to see that no direct limits on $m(P^\pm)$ can be obtained at present for free $m(P^{\pm}_8)$, but at the same time, one can simply read out the lower bound on the mass of color octet PGBs: $m(P^{\pm}_8) > 440\;GeV$ for free $m(P^{\pm})$ (assuming $m_t=180\;GeV$), if we simply interpret the $CLEO$ $95\% C.L.$ limits on the ratio $BR(B \rightarrow X_s \gamma)$ as the bounds on the masses of charged PGBs. Of cause, we have not considered the effects of other possible uncertainties, such as that of $\alpha_s(m_Z)$, next-to-leading-log QCD contribution\cite{AJB}, QCD correction from $m_{top}$ to $M_W$\cite{lcd} etc. The inclusion of those additional uncertainties will broaden the border lines between the allowed regions and excluded regions in Fig.3. The limitations drawn from the calculations will be surely weaken, i.e., the lower limit will become $m(P^{\pm}_8) > 400\; GeV$ at $90\%C.L.$ if we include an additional $20\%$ theoretical uncertainties. As a conclusion, the size of contribution to the rare decay of $b\to s \gamma$ from the PGBs strongly depends on the values of the masses of the top quark and the charged PGBs. This is quite different from the SM case. By the comparison of the theoretical prediction with the current data one can derives out the constraints on the masses of the color octet charged PGBs: $m(P^\pm_8) > 400\;GeV$ at $90\%C.L.$ for free $m(P^\pm)$, assuming $m_t=180\;GeV$. \vspace{1cm} \noindent {\bf ACKNOWLEDGMENT} This work was supported in part by the National Natural Science Foundation of China, the funds from Henan Science and Technology Committee and China Postdoctoral Science Foundation. \vspace{1cm}
\section{Introduction} The option of building a photon-photon interaction region at an $e^{+} e^{-}\;$ linear collider is now taken seriously under consideration. Based on the idea of using laser-induced backscattered photons for inducing high-energy photon collisions, a $\gamma \gamma\;$ collider (PLC) gives rise to new physics opportunity\cite{GinzburgNIM}. The issues concerning electroweak physics will be summarized in this talk\cite{Paris}. Since the symmetry-breaking mechanism remains the last open question in the standard model, an important part of the planning at any future collider must be devoted to that. This obviously includes searches for the Higgs particle and the determination of its properties. One unique opportunity for $\gamma \gamma\;$ colliders in that respect is the direct measurement of the $H\gamma \gamma$ coupling. Should the Higgs searches remain fruitless, the study of the longitudinal $W$ sector would give a handle on the symmetry breaking mechanism. Photon colliders, being essentially a $W$ pair factory, could make a useful contribution in that respect. Before going into the heart of the subject and to give a first idea of the possibilities of $\gamma \gamma\;$ colliders, I will present the main characteristics of high-energy $\gamma \gamma\;$ collisions: \begin{itemize} \item Any elementary charged particle, phase-space allowing, can be produced in $\gamma \gamma\;$ collisions with a model-independent predictable cross-section. \item $\gamma \gamma\;$ gives access to the $J_Z=0$ channel, which is chirality suppressed in $e^{+} e^{-}\;$. To test the electroweak symmetry breaking (ESB) mechanism, this means producing the Higgs as a resonance. \item $\gamma \gamma\;$ collisions feature very large cross-sections, which are always larger than in $e^{+} e^{-}\;$ for the same energy and luminosity. \end{itemize} But all is not so bright, for $\gamma \gamma\;$ colliders also have shortcomings. First, the Higgs resonance cannot be so prominent as the Z at LEP since the coupling of neutral scalars to two photons only occurs at the loop level and is suppressed by a factor $\alpha$. Second, $\gamma \gamma\;$ does not have the same energy as $e^{+} e^{-}\;$ (the maximum energy varies between 80-90\%) and the useful luminosity can be smaller than in $e^{+} e^{-}\;$. The latter is true especially if one uses beams optimized for $e^{+} e^{-}\;$ coliders rather then designed specifically for $\gamma \gamma\;$\cite{Telnovshef}. Finally, the photon collider is not monochromatic, although, as was discussed by Telnov, one can tune the parameters of the laser such as to have a nearly monochromatic spectrum near the maximum energy. This is done at the expense of a drop in luminosity. Certainly, one has a great flexibility in choosing the energy, the spectrum and the polarization of the beams. The choice of spectrum will be dictated by the physics one is interested in. For example, a ``peaked" spectrum where much the luminosity is concentrated over a narrow energy band would be most appropriate to study a resonance. A ``broad" spectrum, one with sizeable luminosity over a wide energy range, would correspond to a multi-purpose machine useful for many processes\cite{Paris}. Whatever the spectrum used, a precise knowledge of it is essential to be able to do precision measurements. More efforts in that direction are needed since recent studies have shown that going beyond the ideal spectrum of Ginzburg et al.\cite{GinzburgNIM} could significantly affect the region where the photons carry only a small fraction of the initial beam energy. This region is particularly sensitive to multiple scattering and nonlinear effects\cite{Chenshef}. For lack of a more realistic spectrum, most of the results presented here use the ideal one. \section{Typical electroweak cross-sections} \renewcommand{\thesection.\arabic{equation}}{\Alph{section}-\arabic{equation}} \setcounter{equation}{0} A comparison of a few characteristic cross-sections in $\gamma \gamma\;$ with the corresponding same final-state processes at $e^{+} e^{-}\;$ clearly show the advantage of the former. Indeed, independent of the spin of the particle and at the same centre-of-mass energy, $\gamma \gamma\;$-initiated processes are, at high enough energy, about an order of magnitude larger than the corresponding $e^{+} e^{-}\;$ reactions (Fig.~1). Even the reaction $\gamma \gamma \rightarrow ZZ$ \cite{Jikiazz,DicusKao}, which is purely a loop effect, rapidly overtakes the corresponding tree-level $e^{+} e^{-}\;$ process. This is due to the rescattering effect $\gamma \gamma \rightarrow W^+W^- \rightarrow ZZ$. Vector-boson production dominates in $\gamma \gamma\;$ collisions due to the t-channel spin-1 exchange. Most prominent is the $W$ pair cross-section, which very quickly reaches a plateau of almost $90$pb. This process is so important that it triggers a host of higher order processes like triple vector production ($\approx 1$pb), 4 vector production ($\approx 100 fb$) or $H$ production via $WWH$ \cite{nousgg3v}. \begin{figure*}[htb] \caption{\label{allgamegam} {\em Typical sizes of non-hadronic $\gamma \gamma\;$, $e\gamma$ and $e^{+} e^{-}\;$ processes. The subscripts in Higgs(top) processes refer to the mass of the Higgs(top). }} \begin{center} \vspace*{.5cm} \vspace*{-1.cm} \end{center} \end{figure*} Large cross-sections are great, but more is required to do interesting physics. A drawback of $\gamma \gamma\;$ processes, especially as regards ESB tests is the small fraction of longitudinal vs transverse $W$'s. While there is a large sample of $W_LW_L$ (in fact more then 5 times than in $e^+e^-$), the extraction of these longitudinals from the transverse background is an arduous task. Since most transverse W's are produced quite forward, imposing angular cuts improves the situation significantly (see Fig.~2). Nevertheless, the ratio LL/TT remains higher in $e^{+} e^{-}\;$. The problem just alluded to is in fact characteristic of all processes that will be discussed for ESB tests in $\gamma \gamma\;$: the extraction of a signal, usually in the longitudinal sector, from a large transverse background. \begin{figure*}[hbt] \vspace*{-.5cm} \begin{center} \caption{\label{eeggwwfig2}{\em Comparing the total $WW$ cross-sections and the longitudinal $W_L W_L$ in $e^{+} e^{-}\;$ {\it vs} $\gamma \gamma\;$. }} \vspace*{-1.cm} \end{center} \end{figure*} \section{Electroweak symmetry breaking} The various options for symmetry breaking can be divided into two classes: light Higgs ($m_H\leq 800 GeV$) or no Higgs (for the discussion here this is equivalent to a heavy Higgs). The implications for electroweak physics differ markedly according to the option one is willing to consider. A light Higgs probably means the existence of supersymmetry unless one is ready to give up the naturality argument raised to avoid the large fine-tuning necessary for the elementary scalar to remain light to all orders. In this option the physics issues at a collider, in particular the PLC, would be the search for the Higgs and measurement of its properties, the search for other supersymmetric particles and determination of their properties, and precision measurements in order to see indirect effects of new physics. An example of the latter is the measurement of the trilinear couplings, $WW\gamma$. These topics can be covered at moderate energies linear colliders ($\sqrt{s_{ee}}=300-500 GeV$) If the light Higgs does not exist, then ESB is triggered by strong forces, the scale being set by $\Lambda=4\pi v\approx 1$~TeV. Although the details of the model are not known there must be new physics at this scale (e.g. technicolour, strongly-interacting particles). In particular, this new physics would show up in $W$ self-interactions or in $W_LW_L$ scattering. The connection between heavy Higgs and longitudinal W's is best established via the process $W_L W_L\rightarrow W_L W_L$. In the SM the Higgs is introduced to cure the bad high-energy behaviour of this amplitude; nonetheless $W_LW_L$ interactions become strong if $m_H\approx .8-1$~TeV. The physics of the heavy Higgs is most relevant for TeV linear colliders. Since no model has gained a consensus to describe the strongly-interacting electroweak sector, one must strive for a model-independent description of this sector. One approach, which is valid up to some scale $\Lambda$, uses an effective chiral Lagrangian. Assuming a custodial $SU(2)$ symmetry to ensure that the parameter $\rho\approx 1$, new physics in the weak boson sector is described by nonrenormalizable terms suppressed by powers of $1/\Lambda$, \begin{eqnarray} {\cal L}={\cal L}_{SM} (no Higgs)+\sum \frac{1}{\Lambda^n} {\cal L}_n \end{eqnarray} \noindent The leading order chiral Lagrangian, ${\cal L}_2$, contains only the mass terms for the vector bosons while the Next-to-Leading order, ${\cal L}_4$, contains the self-interactions. This includes trilinear or quartic interactions of massive vector bosons and at most one photon\cite{Hawai}. Self-interactions including two photons only appear at higher order, ${\cal L}_6$. The effective Lagrangian formalism would break down if the scale at which the experiment is performed is sufficient to produce new resonances. These must be explicitly incorporated, the cases of either a scalar, vector or tensor resonance will be considered. The scalar one ($\sigma$-like) is representative of a heavy Higgs while the vector one ($\rho$-like) occurs in technicolour. To cover all standard and non-standard manifestations of symmetry breaking, the strategy at the future collider must include: tests of $W$ self-interactions and of longitudinal vector-boson scattering, the search for the Higgs in the whole range of possible masses, as well as searches for other heavy resonances. All these aspects can be tackled at a photon collider, as will be described in the rest of this talk. \section{$\gamma \gamma \ra W^+ W^-\;$ and W self-interactions} The importance of this process cannot be over-emphasized considering the large cross-section involved. Although the bulk of the reaction is due to the gauge transverse sector, the fact that there are so many $W$'s around makes this reaction the ideal place to conduct precision tests of the electromagnetic couplings of the $W$. In the effective chiral Lagrangian description of Higgsless models, the anomalous trilinear couplings invoke two C and P conserving operators at Next-to-Leading order, $L_{9L}$ and $L_{9R}$, and one C and P violating operator $L_{C}$\cite{Hawai}. The latter affects only $ZWW$ and $\gamma ZWW$ interactions while only the combination $L_{9L}+L_{9R}$ contribute to $\gamma WW$. The $L_i$ operators are expected to be ${\cal O} (1)$. There is an extensive literature \cite{Paris} on the effect of anomalous couplings in various experiments. Comparisons of different analyses have shown that a $500$~GeV linear collider with a luminosity of $10 fb^{-1}$ does significantly better then the LHC for trilinear couplings. Furthermore, the limits that can be obtained in $\gamma \gamma\;$ at this energy, $ |L_{9L}+L_{9R}|< 10 $ represent a 50\% improvement over $e^{+} e^{-}\;$\cite{Hawai}. This is shown in Fig.~3. However, this result is not sufficient to reach the level where one expects new physics to set in. In $e^{+} e^{-}\;$ it was shown recently\cite{steve,barklow} that meaningful limits could be obtained with a luminosity ${\cal L}=50-80 fb^{-1}$. It remains to be seen if the same can be done at a $\gamma \gamma\;$ collider. For that one needs to generate the four-fermion final state from the decay of the $W$'s, while keeping the full spin-correlation. \begin{figure*}[htb] \caption{\label{gg19}{\em Comparison between the expected bounds on the two-parameter space $(L_{9L},L_{9R})$ at the NLC500, LHC and LEP2. We also show (``bars") the limits from a single parameter fit.}} \begin{center} \vspace*{-1.cm} \vspace*{-1.cm} \end{center} \end{figure*} A $\gamma \gamma\;$ collider can do more than precision tests on $\gamma WW$. It is also sensitive to $ZWW$ couplings through processes with three particles in the final state, for example $\gamma \gamma \rightarrow WWZ$\cite{nousggwwz}, $e\gamma \rightarrow e WW$ and $e\gamma\rightarrow \nu W Z$. The latter is very sensitive to the operator $L_{C}$\cite{Dawsong5}. The limits obtained, $L_{C}<25$ at 500 GeV are comparable to the ones from $e^+e^-\rightarrow W^+W^-$\cite{BMT}. Furthermore, the sensitivity increases rapidly with energy. \subsection{Effect of radiative corrections} When doing precision tests one must worry about the effect of radiative corrections that could mimic those of the new couplings. Recently the complete one-loop SM corrections for helicity amplitudes for $\gamma \gamma \ra W^+ W^-\;$ were calculated\cite{Dennerradcor}. It turns out that the radiative corrections for this process are theoretically clean due to the absence of most universal leading corrections. The running of $\alpha$ is irrelevant since we are dealing with on-shell photons, all uncertainties due to $\log(m_q^2)$ terms in small masses disappear, and there are no large log corrections associated with colinear photons except at very high energies. Furthermore, the corrections are not very sensitive to either $M_{t}$ or $M_{H}$ except near the resonance. Although some helicity amplitudes receive huge corrections, they are precisely the ones that contribute very little to the total cross-section. Typically radiative corrections between $1-10\%$ at $\sqrt{s_{ee}}=500$~GeV are obtained, and they tend to increase with energy ($\approx 20\%$ at $1$~TeV). In any case, the inclusion of radiative corrections are not expected to change much the previously obtained results on measurements of trilinear couplings. Considering the large numbers of $W$'s available, there are other interesting questions that can be studied in $\gamma \gamma\;$ which I have not addressed here. Among them are the possibility of direct tests of quartic couplings involving photons\cite{nousggvv}, CP tests in $W$ decay and measurement of the $Wtb$ coupling which could also give some clues about symmetry breaking. \section{Higgs searches} One of the most attractive motivations for doing physics with very energetic photon beams is the unique capability of this mode for producing a scalar particle, such as the Higgs, as a resonance. I have already mentioned that the coupling of the Higgs to two photons occurs only at the loop level. It should be emphasized that a precision measurement of the $H\gamma \gamma$ coupling is an indirect way of revealing all massive charged particles that could be present in an extension of the standard model. These heavy quanta would not decouple and would contribute to the production rate in $\gamma \gamma\;$. While many processes are sensitive to the presence of the Higgs (see Table 1), the prime interest of the photon mode lies in the Intermediate Mass region \footnote{$\gamma \gamma\;$ is also very useful in the mass range below 90 GeV, a case can be made for building a low-energy dedicated $\gamma \gamma\;$ collider in the event of a discovery of the Higgs at LEP2.}. For such a Higgs, the main decay mode is into $b\overline{b}$. Although a search is feasible at LHC, it will be a difficult and long task to extract a signal in this case. For heavier Higgs masses the resonance can be seen in the $WW$ \cite{Ginzburgshef} or $ZZ$ channel. However, the usefulness of these modes is tamed by the presence of large backgrounds from transverse vector bosons. Ultimately, for a Higgs above 400 GeV, and regardless of the energy available for the PLC, one would have to resort to other channels such as associated Higgs production or $WW$ fusion (via the process $\gamma \gamma\rightarrow WWWW$). \baselineskip=12pt \begin{table*}[htb] \caption{\label{Higgsmass} {\em Processses for Higgs searches at PLC and other colliders}} \vspace*{0.3cm} \centering \begin{tabular}{|l|l|l|l|} \hline Mass&Collider& PLC&$\sqrt{s_{ee}}$ \\ \hline &&&\\ $M_{H}<65 GeV$& LEP& Ruled out&-----\\ $65 GeV<M_{H}<90GeV$&LEP2&$\gamma \gamma\rightarrow H\rightarrowb\overline{b}$&$.1-.5$~TeV\\ $90 GeV<M_{H}<140GeV$&NLC & $\gamma \gamma\rightarrow H\rightarrowb\overline{b}$&$.2-.5$~TeV \\ $140 GeV<M_{H}<200GeV$&LHC &$\gamma \gamma\rightarrow H\rightarrow WW$&$.5$~TeV \\ $200 GeV<M_{H}<400GeV$&LHC &$\gamma \gamma\rightarrow H\rightarrow ZZ$&$1$~TeV \\ $400 GeV<M_{H}<700GeV$&LHC & $\gamma \gamma\rightarrow WWWW$&$2$~TeV\\ \hline \end{tabular} \end{table*} \baselineskip=14pt \subsection{Intermediate Mass Higgs} For the IMH, $\gamma \gamma\;$ can contribute in the discovery mode or perform precision measurements of its properties. A crucial point relates to the choice of the spectrum and polarization used. Since the Higgs is produced only in the $J_z=0$ channel, polarization plays a crucial role in enhancing the signal over background. Assuming the Higgs has been found and its mass measured, one could tune the energy of the collider and the parameters of the laser such that the peak of luminosity lies precisely at $\sqrt{s_{\gamma \gamma}}=M_{H}$. This is obviously the preferred way to operate when measuring $H\gamma \gamma$, though one has to realise that good luminosity is required. Early estimates for a 500 GeV collider and a luminosity of ${\cal L}=20 fb^{-1}$ give a 10\% precision on the width,\cite{Borden} the effect of background from one radiated gluon is discussed by Khoze \cite{Khozeshef}. One disadvantage of operating in that mode is that the PLC would be run at energies much below the nominal $e^{+} e^{-}\;$ energy, precluding the study of interesting processes such as the $W$ pair production and other $W$ reactions that could occur at higher energy. Furthermore, this low-energy narrow-band scheme could render kinematically inaccessible some of the particles that would only be probed indirectly in $H\gamma \gamma$, not to mention that the $\gamma \gamma\;$ mode, when operated in the full range of energy, can access scalar particles that would be kinematically out of reach in the $e^{+} e^{-}\;$ mode. If one would have two interaction regions, one devoted to $\gamma \gamma\;$ the other to $e^{+} e^{-}\;$, and if the Higgs has not been found elsewhere, $\gamma \gamma\;$ could be used to search for the Higgs. The method that allows for simultaneous studies of processes at high energy consists of running the PLC using a ``broad" spectrum so that one would have reasonable luminosity over a range of energies. The main problem in the Higgs searches, whatever the scheme used, lies in the large background. The prominent one comes from direct QED $\gamma \gamma \rightarrow q \bar q$ production where $q=b$ or other light quark flavours, in particular charm. However this background can be dealt with since the bulk of the cross-section is in the forward direction, so that a modest angular cut could efficiently suppressed this background and would almost totally eliminate its $J_Z=0$ contribution. Therefore a spectrum with a predominantly $J_Z=0$ component would both enhanced the signal and reduce the background. When the PLC is run in the ``broad" spectrum mode there are other more important backgrounds that have to be taken into account\cite{Halzen,Higgsreson}. They arise from the hadronic structure of the photon which can resolve into a gluon or a quark with some spectator jets left over. One then has to worry about $q \bar q$ production through $\gamma g$ as well as a host of 1-resolved and 2-resolved process. These backgrounds dominate the signal. However, since most of the resolved events are very boosted, judicious cuts can reduce it to a manageable level. It was shown, using the ideal spectrum, that at $500$~GeV with ${\cal L}=10 fb^{-1}$, one could obtain a good signal for $M_H=110-140$~GeV \cite{Higgsreson}. Furthermore, the situation improves for a collider of lesser energy, due to the reduced resolved background. For example, at $350$~GeV a signal is easily extracted for the whole IMH range. Of course this assumes the ideal spectrum. However, for the masses considered, the signal falls in the region where the spectrum is most severely affected by effects of multiple scattering and nonlinear effects. These questions should be reassessed taking these effects into account as the conclusions could differ drastically. There have been suggestions to determine directly the parity of the Higgs using linear polarizations of the photon\cite{Gunionparity}. Since the degree of linear polarization is never very large ($<30\%$), this always requires large luminosities, ${\cal L}=100 fb^{-1}$. \subsection{Associated production} For the IMH, it will be hard to unravel a peak formation if the collider energy is greater than $500$~GeV. As will be discussed in the next section, the resonance will remain hidden for heavier Higgs ($M_{H}\ge 400$~GeV) even if one uses the most favourable channel, $ZZ$. Fortunately, other efficient mechanisms for Higgs production are available, in particular the radiation of a Higgs from a $W$ pair. This is to be expected since the cross section for $W$ pair production is so large and the Higgs couples preferentially to the weak bosons. In fact, at 1 TeV, before folding with the luminosity spectrum, the $e\gamma$, $\gamma \gamma\;$ and Bjorken process are comparable for all Higgs masses\cite{nousgg3v}. However, the $\gamma \gamma\;$ production mode is suppressed when including a more realistic photon luminosity spectrum. Still, at $1$~TeV one obtains a measurable cross-section ($\sigma >3fb$) for $M_{H}<400 GeV$. \subsection{\boldmath{ $\gamma \gamma \rightarrow ZZ$}} At first this reaction was believed to provide a background-free environment for either Higgs production or non-standard physics signals in $Z_LZ_L$ since it is purely a loop process in the ${\cal{S}} {\cal{M}}\;$. The first full calculation by Jikia\cite{Jikiazz} of the one-loop process $\gamma \gamma \ra ZZ \;$ within the ${\cal{S}} {\cal{M}}\;$ dampened this enthusiasm since it turned out that, once again, the transverse modes are overly dominant, especially at high energy. This is due essentially to the $W$ loops, the $WW$ produced in $\gamma \gamma\;$ rescatter into $ZZ$. At $\sqrt{s_{ee}}=400$~GeV, the Higgs resonance is clearly evident over the $TT$ continuum all the way up to the kinematic limit. With $\sqrt{s_{ee}}=500$~GeV, it already becomes difficult to extract a Higgs with $M_H\sim350$~GeV. \cite{Jikiazz} To obtain these results, Jikia used a predominantly $J_Z=0$ spectrum that is peaked towards the maximum $\hat{s}_{\gamma \gamma}$, this is not the optimum choice. With a broader spectrum featuring a dominant $J_Z=0$ for small $M_{ZZ}$, one could still see a peak in the $M_{ZZ}$ invariant mass for Higgs masses up to $400$~GeV at a $1$~TeV $e^{+} e^{-}\;$ machine \cite{DicusKao}. {}From the perspective of observing the Higgs resonance beyond TeV $e^{+} e^{-}\;$ energies, the situation becomes totally hopeless as the transverse $ZZ$ are awesome\cite{Jikiazz}. \section{Strongly-interacting electroweak sector (SEWS)} If the Higgs is not found at LHC, or in the sub-TeV version of NLC, we will be in the realm of the SEWS. This sector would be probed most efficiently at TeV energies through the reaction $V_L V_L\rightarrow V_L V_L $ ($V=W$ or $Z$). In this channel one would either search for a resonance or, if the energy is not sufficient, for new interactions such as the ones described by the effective chiral Lagrangian. The $V$ pair-production processes could be regarded as the testing ground for possible rescattering effects in $WW\rightarrow VV$ that originate from the symmetry breaking sector. Unfortunately, at high energies, {\it i.e.}, at high $VV$ invariant masses, where the effect of the New Physics would be most evident, one has to fight extremely hard against the background for transverse $W$ and $Z$. Indeed, recent analyses have shown that while it might be possible to see effect of a tensor resonance, a scalar one as well as indirect effects are hopeless\cite{BergerBerkeley}. One then has to resort to the only source of longitudinal vector bosons, the ones taking part in the fusion process and contributing to $WWWW$ or $WWZZ$ production. This process is the analog of $e^+e^-\rightarrow \nu\overline\nu W^+W^-$ and was originally believed to be more favourable due to a presumed larger $W_L$ content in the photon than in the electron. While it is true that in the photon there is an additional structure function corresponding to the spectator $W_L$, the dominant contribution is from transverse spectator $W$'s. The latter features basically the same structure function as in the electron, except for an overall factor\cite{Paris}. One would therefore expect that $\gamma \gamma\;$ should be comparable to $e^{+} e^{-}\;$ at the same energy and luminosity. This conclusion was born out by two independent exact calculations of this ${\cal{S}} {\cal{M}}\;$ process \cite{Jikiawwww,Cheungwwww}. The signal of a heavy Higgs is a significant increase in the channels with at least three $W_L$. To extract a signal requires tagging all four $W$'s, the spectator ones being associated with the low $p_T$ and the longitudinal ones with the central $W$'s. The spectators are tagged with one hadronic and one leptonic decay while the central ones go into four jets. The results of the analysis showed that a 2~TeV PLC (${\cal L}=10 fb^{-1}$) would give a good signal ($S/\sqrt{B}\approx 10$) for a heavy Higgs-like scalar of 1TeV\cite{Cheungwwww}. This is comparable to the $e^{+} e^{-}\;$ process. However, the inclusion of the spectrum has a dramatic effect and a linear collider of $2$~TeV in the $e^{+} e^{-}\;$ center of mass with ${\cal L}=200fb^{-1}$ is needed to reach the same significance level. Another interesting conclusion from these calculations is that a signal for a Higgs of 400-700~GeV can easily be seen with $\sqrt{s_{ee}}=1.5$~TeV and ${\cal L}=200 fb^{-1}$. The PLC can therefore cover the whole mass range for light or heavy Higgs searches provided a good choice of energy and spectrum is made, although precision measurements are possible only for light Higgs ($M_{H}< 120 GeV$). \section{Search for new particles: supersymmetry} As the best motivated alternative to the standard model, one should investigate the consequence of supersymmetic models. Supersymmetry would provide a natural framework for light Higgses. The three neutral scalars of supersymmetric models, $h$,$H$,$A$ (pseudoscalar) could be produced as a resonance in $\gamma \gamma\;$. This reaction would then extend the reach in mass of $e^{+} e^{-}\;$ since in the latter $H$ and $A$ can only be produced together and require $\sqrt{s_{ee}}>M_H+M_A$. At $\sqrt{s_{ee}}=500$~GeV, using the $b\overline{b}$ mode, this gives the following discovery region for the supersymmetric scalars\cite{GunionHaber}: $ 110$~GeV$< M_{H} < 200$~GeV and $100$~GeV$<M_A < 2 M_{t}$. Recently, it was pointed out that this was true only if scalars decayed primarily via ${\cal{S}} {\cal{M}}\;$ final states. Otherwise the above limits require high luminosities ${\cal L}> 60 fb^{-1}$\cite{Gunionsusy}. The $\gamma \gamma\;$ collider can search also for other supersymmetric particles \cite{Murayama},the production cross-sections being universal, were already shown. Typically, one finds that $\gamma \gamma\;$ can have good cross-sections but offer little advantage over the $e^{+} e^{-}\;$ mode, in part because of the lower achievable energy. It is for selectron searches in $e\gamma\rightarrow \tilde{e}\tilde{\chi}$ that the laser scheme becomes extremely useful, as the discovery of a selectron of $m_{\tilde{e}}\approx\sqrt{s_{e\gamma}}$ is possible. \section{Conclusion} A $\gamma \gamma\;$ collider of energy ranging from .2 to 2~TeV should prove to be a useful tool for probing the electroweak symmetry-breaking sector through either Higgs searches or $W$ physics. It is unique in producing a scalar on resonance and is complementary to an $e^{+} e^{-}\;$ collider in many processes. \noindent {\bf {\large Acknowledgements}} I am most grateful to my friends and collaborators Marc Baillargeon and Fawzi Boudjema for all the work on electroweak physics issues. I also thank George Jikia for kindly supplying the curves for $\gamma \gamma\rightarrow WWWW$. \vspace{.5in}
\section{INTRODUCTION} Spectroscopy in the superdeformed minimum has reached a certain level of maturity to justify a phenomenological analysis of the available data (see \cite{jan91,nol94} for recent reviews). Such an approach would be useful in systematizing the data and would also provide a complimentary perspective to the more microscopic theories. For this purpose, we use the interacting boson model (IBM) \cite{iac87} which has been established as one of the simplest and most versatile collective models. It has been especially successful in correlating spectroscopic data in deformed nuclei in terms of a few parameters of a quadrupole Hamiltonian \cite{cas88}. Microscopic study of the nucleon pair structure in the superdeformed well \cite{hon92} indicates that, compared to the deformed nuclei, they have about three times more active pairs of nucleons (bosons), and the $L=4$ pair ($g$ boson) plays a much more significant role. As numerical diagonalization of an $sdg$-IBM Hamiltonian is not possible for more than 10 bosons, one needs alternative methods of solution to apply the IBM to superdeformed nuclei. Here, we use the angular momentum projected mean field theory which leads to a $1/N$ expansion for all matrix elements of interest \cite{kuy88}. Accurate representation of high-spin states in the $1/N$ expansion formalism requires terms to order $(L/N)^6$ which have been obtained recently using computer algebra \cite{kuy95}. The extended formalism provides an analytical method for analysis of superdeformed states which is both accurate and efficient. The plan of the paper is as follows. After reviewing the microscopic basis of the IBM for superdeformed nuclei, we introduce the $1/N$ expansion formalism and discuss its recent extension to higher orders. We then use the $1/N$ expansion formulas for a quadrupole Hamiltonian to study systematic features of dynamic moment of inertia and $B(E2)$ values. The results are used in a description the superdeformed bands in the Hg-Pb and Gd-Dy regions. \section{MICROSCOPIC BASIS} In this section, we study the typical structure of strongly deformed states and investigate the relation between the superdeformation and the IBM based on \cite{hon92}. For this purpose, we use the Nilsson + BCS model with particle number projection. Superdeformed states can be characterized as ground states in a superdeformed potential well which is separated from the normal one by a potential barrier. For such ground-like states which show strongly collective nature, this model seems to work well. Using the experimental deformation parameters and electric transition probabilities (or moments) as input, one can obtain reasonable wave functions. These wave functions are analyzed from the viewpoint of collective nucleon pairs, which leads to a natural extension of the usual IBM. We briefly summarize the formulation of the Nilsson + particle-number-conserving BCS model. The single particle orbits in a deformed potential are described well by the Nilsson Hamiltonian \cite{nil55} \begin{eqnarray} H_{\rm Nilsson} &=& -\frac{\hbar^{2}}{2m} \nabla ^2 +\frac{m}{2}\omega_{0}^{2}r^{2} [1-\frac{4}{3}\delta P_{2}(\cos\theta)] \nonumber \\ &-& 2\hbar \omega_{0} \kappa \mbox{\boldmath $l$} \cdot \mbox{\boldmath $s$} -\hbar \omega_{0} \kappa \mu (\mbox{\boldmath $l$}^{2}-\langle \mbox{\boldmath $l$}^{2} \rangle_{N}), \end{eqnarray} where $\delta$ is the deformation parameter and $P_{2}(\cos\theta)$ denotes the Legendre polynomial. The term $\langle \mbox{\boldmath $l$}^{2} \rangle_{N} = \frac{1}{2}N(N+3)$ is the expectation value of $\mbox{\boldmath $l$}^{2}$ averaged over one major shell with the principal quantum number $N= 2n+l$. The value of the oscillator frequency for a mass-A nucleus is determined from $\hbar\omega_{0} = 41A^{-\frac{1}{3}}$, which is about 7.1 MeV for the superdeformed nuclei in the Hg-Pb region ($A \sim$ 190) and 7.7MeV for those in the Dy-Gd region ($A \sim$150). We use the usual values for the parameters $\kappa$ and $\mu$ which are 0.0637, 0.60 for proton orbits and 0.0637, 0.42 for neutron orbits, respectively \cite{gus67}. In order to include short range correlations, the monopole pairing interaction is added to the Nilsson hamiltonian \cite{nil61} \begin{equation} H = H_{\rm Nilsson} + G P^{\dagger} P, \end{equation} where $G$ denotes the pairing strength parameter, and \begin{equation} P^{\dagger} = \sum_{k>0} c_{k}^{\dagger} c_{\bar k}^{\dagger} \ , \end{equation} is a pair creation operator. Here $c_{k}^{\dagger}$ stands for the creation operator of a nucleon in the spherical single particle orbit $k$, and $\bar k$ denotes the time reversed state of $k$. This Hamiltonian is solved by the variation using a BCS wave function \begin{equation} \mid \Psi \rangle = \prod_{\alpha>0} (u_{\alpha} + v_{\alpha} a_{\alpha}^{\dagger} a_{\bar \alpha}^{\dagger}) \mid 0 \rangle, \end{equation} where $a_{\alpha}^{\dagger}$ denotes the creation operator for a nucleon in the deformed canonical (Nilsson) orbit labeled by $\alpha$. The particle number conservation has been found to be important in the case of weak pairing correlations and also for moments of inertia of high spin states in the cranking calculation of the superdeformed states \cite{dud88,shi90}. Thus we carry out the particle number projection before variation according to the method given in Ref. \cite{egi82}. The solution corresponds to the minimum of the number projected energy \begin{equation} E^{P}[\Psi] = \frac{\langle \Psi \mid H P^{N} \mid \Psi \rangle} {\langle \Psi \mid P^{N} \mid \Psi \rangle}, \end{equation} where $P^{N}$ denotes the particle number projection operator. The deformation parameters of the superdeformed states in the Hg-Pb (Dy-Gd) region are given by $\delta \sim$ 0.40 (0.50), which is equivalent to the axis ratio of 5:3 (2:1). Because of this strong deformation, it is insufficient to take only one active major shell and take into account the corrections due to the core-polarization effect through renormalization in one major shell. Thus we first seek a suitable model space for description of superdeformed states. For simplicity, we turn off the pairing force which is not important for this purpose. We take $^{194}$Hg ($N$=114, $Z$=80) and $^{152}$Dy ($N$=86, $Z$=66) as examples of the Hg-Pb and the Dy-Gd regions, respectively. In order to define the model space necessary for description of superdeformed states, we utilize the intrinsic quadrupole moment $Q_0$. The intrinsic quadrupole moment is calculated in the space of all spherical orbits up to the principal quantum number $N=N_{\rm max}$. Then $N_{\rm max}$ is increased until the value of $Q_0$ is saturated to a good extent. {}From this procedure we obtain $N_{\rm max}=12$. The corresponding values of $Q_0$ for proton and neutron orbits are 19{\it b} and 29{\it b} for $^{194}$Hg, and 19{\it b} and 27{\it b} for $^{152}$Dy, respectively. Note that the experimental values of $Q_0$ are $18\pm 3~eb$ for $^{152}$Dy \cite{ben91} and $17\pm 2~eb$ for $^{194}$Hg \cite{hug94}, which are consistent with the present results if we take the bare charges, $e_{\rm p}=1$ and $e_{\rm n}=0$. We now consider the inert core of superdeformed states. The Nilsson wave function is obtained by putting all nucleons in the Nilsson orbits from the bottom. One Nilsson orbit can be expanded as a linear combination of many spherical harmonic oscillator orbits, and the square of expansion coefficients gives the occupation probability of each spherical orbit. We expand all the occupied Nilsson orbits and sum up all the occupation probabilities which belong to the same spherical orbits, to obtain the total occupation probability for a given spherical harmonic oscillator basis. Due to the strong quadrupole field, one Nilsson orbit spreads over many spherical orbits. Thus the orbits with very high single particle energy can gain some finite occupation probabilities, while the occupation of the orbits with small single particle energy may become incomplete. Nevertheless several lower spherical orbits are occupied almost completely and can be considered as a new inert spherical core for the superdeformed states. Note that we do not take the usual ``hole'' picture as it is meaningful only for states whose configuration are well described within one major shell. First consider the case of $^{194}$Hg. In Fig.~\ref{fig1} the occupation probability of each spherical harmonic oscillator orbit is shown for neutrons (a) and protons (b). The orbits are ordered according to their single particle energy at $\delta=0$ as $1s_{1/2}$, $1p_{3/2}$, $\cdots$. The case of $\delta=-0.13$ which simulates the deformation of normal oblate states is also shown for comparison. For normal deformation, it is seen from Fig.~\ref{fig1}-a that the occupation of the proton orbits is almost complete at $2d_{5/2}$ ($Z=64$), while the occupation probability is almost vanished for orbits above $Z=82$. These results suggest that we can consider the $Z=64$ subshell as an inert core and three valence orbits ($2d_{3/2}$, $3s_{1/2}$ and $1h_{11/2}$) as active. This gives the valence proton number as $Z_{v}=16$. In the case of $\delta=0.40$, the proton orbits are almost completely occupied up to $Z=50$. Above $Z=50$, the occupation probability drops suddenly though it remains about 10$\%$ over many orbits. Clearly, one should incorporate the contributions of these high energy orbits. Thus it is reasonable to take the $Z=50$ spherical inert core and include quite many orbits above there as active valence orbits. In this case the valence proton number becomes $Z_{v}=30$. In the same way, it can be seen form Fig.~\ref{fig1}-b that the spherical inert core for neutron orbits are $N=100$ ($N_{v}=14$) and 82 ($N_{v}=32$) for $\delta=-0.13$ and 0.40, respectively. The active valence orbits for $\delta=-0.13$ are $2f_{5/2}$, $3p_{3/2}$, $3p_{1/2}$ and $1i_{13/2}$, while for $\delta=0.40$ it is still insufficient to include only two or three major shells. We can see a similar behaviour of occupation probabilities in the wavefunctions of $^{152}$Dy, which is shown in Figs. \ref{fig1}-c for proton and \ref{fig1}-d for neutron orbits. In these figures two cases of $\delta=0.50$ (superdeformed state) and 0.25 (normal prolate state) are compared. It is clear that $Z=50$ and $N=82$ inert cores are good for normal states, while $Z=40$ and $N=50$ cores are suitable for superdeformed states. Thus the inert core of superdeformed states becomes much smaller than that of normal states in both the Hg-Pb and Dy-Gd regions. {}From the viewpoint of the IBM, the number of bosons is determined by half of the number of valence nucleons. Because of the small inert core the number of bosons increases significantly for superdeformed states in comparison with that in the usual IBM. In fact in the case of $^{194}$Hg, the boson number in the usual IBM is $N_{\rm normal}=(82-80)/2+(128-114)/2=8$ by taking the usual hole picture, and $N_{\rm normal}=(80-64)/2+(114-100)/2=15$ with the particle picture mentioned above. On the other hand, the number of bosons for superdeformation becomes $N_{\rm super}=(80-50)/2+(114-82)/2=31$. Similarly, in the case of $^{152}$Dy, $N_{\rm normal} = (66-50)/2 + (86-82)/2 = 10$ while $N_{\rm super} = (66-40)/2 + (86-50)/2 = 31$. In general, $N_{\rm super}$ is about three times larger than $N_{\rm normal}$. Note that the number of proton bosons and neutron bosons are close in these two cases, and this approximate equality seems to be a general tendency of the superdeformed states. This result can be naturally understood since equal numbers of valence protons and neutrons maximizes the attractive proton-neutron interaction. Next we consider the effects of pairing correlations on the structure of wave functions of superdeformed states. The strength parameter $G$ of the pairing interaction should be chosen depending on the model space. Since the value of $G$ for such a large space is not known empirically, we first describe normal states within the extended valence space, and determine the value of $G$ by requiring that the pairing gap ${\it \Delta}$ takes a reasonable value. For $^{194}$Hg the value of $G$ has turned out to be 0.06MeV which gives ${\it \Delta}\sim 1$ MeV. Using this value, we investigate the effect of the pairing correlations on the structure of valence wave functions. For this value of $G$, the gap for superdeformed states becomes about $\Delta =0.5$ MeV for both proton and neutron orbits. In contrast to normal deformed states, which are sensitive to changes in values of $G$, the superdeformed states are almost insensitive to $G$ values (the intrinsic quadrupole moment and the occupation probabilities change very little). Thus the following discussion about the structure of valence wave function of superdeformation is almost independent of pairing correlations. We can investigate the relation between the superdeformation and the IBM by analyzing valence wave functions from the viewpoint of collective nucleon pairs. Since the bosons in the IBM are understood as images of these pairs, such an analysis is essential in establishing a microscopic basis for the super IBM. We consider $^{194}$Hg as an example. The Nilsson + particle-number-conserving BCS wave function can be expressed as the condensed state of coherent Cooper-pairs in the deformed potential \cite{ots86} \begin{equation} P^{N} \mid \Psi \rangle \propto ({\it \Lambda}_{\pi}^{\dagger})^{N_{\pi}} ({\it \Lambda}_{\nu}^{\dagger})^{N_{\nu}} |0 \rangle , \end{equation} acting on the inert core $|0 \rangle$. In this expression, ${\it \Lambda}_{\pi}^{\dagger}$ (${\it \Lambda}_{\nu}^{\dagger}$) denotes the creation operator of a Cooper-pair in proton (neutron) orbits and $N_{\pi}$ ($N_{\nu}$) means half of the valence proton (neutron) number. These ${\it \Lambda}$-pairs can be decomposed into a linear combination of collective nucleon pairs with good angular momenta \begin{equation} {\it \Lambda}^{\dagger} = x_{0}S^{\dagger} + x_{2}D_{0}^{\dagger} + x_{4}G_{0}^{\dagger}+\cdots , \end{equation} where $S^{\dagger}$, $D^{\dagger}$, $G^{\dagger}$, $\cdots$ denote the collective nucleon pairs with spin-parity $J^{\pi}=$ $0^{+}$, $2^{+}$, $4^{+}$, $\cdots$ and the $x_{J}$'s are amplitudes. The probability of each pair in the ${\it \Lambda}$-pair is given by the square of each amplitude, and is listed in Table~\ref{table1} for two cases of $\delta=-0.13$ and $\delta=0.40$. It is well known that in the case of normal deformation the dominant components are the $S$- and $D$-pairs \cite{ots82,bes82}. In fact, these two components account for 100$\%$ probability in the case of $\delta=-0.13$. In the case of $\delta=0.40$, the total probability of the $S$- and $D$-pairs is about 80$\%$ and we can conclude that these pairs are still dominant in the ${\it \Lambda}$-pair. However the probability of the $G$-pair is now sizable, and it can no longer be neglected in a detailed description of high-spin states. It should be noted that the ratio of the $S$-pair to the other pairs is quite similar to that of $s$-boson to the other bosons in the SU(3) limit of the IBM, which are shown in the same table. This suggests that the SU(3) limit of the $sdg$-IBM could provide a reasonable phenomenological framework for superdeformed states. To summarize the microscopic results, we emphasize two important points for the description of superdeformed bands in the IBM: One is the significant increase in the boson number, and the other is the importance of g-bosons. In addition, it has been found that the bosons for superdeformed states carry the collectivity over many major-shells and that the SU(3) limit is a reasonable starting point. \section{1/N EXPANSION FOR SUPER IBM} A simultaneous description of the spectroscopy of normal and superdeformed states requires rather complicated wave functions, therefore we focus on the latter here and leave the complete picture for future work. We introduce the superbosons ${\bf s}, {\bf d}, {\bf g}$ as the boson images of the $S, D, G$ collective nucleon pairs in the superdeformed well (bold face notation is used for super bosons to distinguish them from the normal ones). The quadrupole Hamiltonian for this system of bosons has the form \begin{equation} H=-\kappa {\bf Q} \cdot {\bf Q}, \label{ham} \end{equation} where the quadrupole operator is defined as \begin{equation} {\bf Q} = [{\bf s}^\dagger {\tilde {\bf d}} + {\bf d}^\dagger {\tilde {\bf s}} ]^{(2)} + q_{22} [{\bf d}^\dagger {\tilde {\bf d}} ]^{(2)} + q_{24} [{\bf d}^\dagger {\tilde {\bf g}} + {\bf g}^\dagger {\tilde {\bf d}} ]^{(2)} + q_{44} [{\bf g}^\dagger {\tilde {\bf g}} ]^{(2)}. \label{qsdg} \end{equation} Here brackets denote tensor coupling of the boson operators and $\tilde b_{lm}=(-1)^{m}b_{l-m}$. The parameters $q_{jl}$ in Eq. (\ref{qsdg}) determine strengths of boson interactions relative to the $s-d$ coupling. Since the SU(3) limit is used as a reference point in the rest of the paper, we quote the values for the quadrupole parameters in this limit; $q_{22}=-1.242$, $q_{24}=1.286$, $q_{44}=-1.589$. As stressed in the introduction, numerical diagonalization of this Hamiltonian for $N \sim 30$ bosons is not possible even on a supercomputer. The large number of bosons are, however, an advantage for the analytic $1/N$ expansion technique which we employ here for solving the Hamiltonian Eq. (\ref{ham}). The $1/N$ expansion method has previously been discussed in detail \cite{kuy88} and the recent extensions to higher orders are given in Ref.~\cite{kuy95}. Therefore, we give only a short account of the formalism here, focusing mainly on the accuracy of the results for high-spin states. The starting point of the $1/N$ calculations is the boson condensate \begin{equation} |N,{\bf x}\rangle =(N!)^{-1/2}({\bf b}^\dagger)^N|0\rangle,\quad {\bf b}^\dagger= x_0 {\bf s}^\dagger + x_2 {\bf d}_0^\dagger + x_4 {\bf g}_0^\dagger, \end{equation} where $x_l$ are the mean field amplitudes to be determined by variation after projection (VAP) from the energy expression \begin{equation} E_L=\langle N,{\bf x}| H P^L_{00} | N,{\bf x}\rangle / \langle N,{\bf x}| P^L_{00} | N,{\bf x}\rangle. \end{equation} Here $P^L_{00}$ denotes the projection operator. The resulting energy expression is a double expansion in $1/N$ and $\bar L=L(L+1)$, and has the generic form \begin{equation} E_L = N^2 \sum_{n,m} {e_{nm}\over (aN)^m} \Bigl({\bar L \over a^2N^2}\Bigr)^n, \label{me1} \end{equation} where $a=\sum_l \bar l x_l^2$ and the expansion coefficients $e_{nm}$ involve various quadratic forms of the mean fields $x_{l}$, e.g., $e_{00}=(\sum_{jl} \langle j0 l0|20 \rangle q_{jl} x_j x_l)^2$. The coefficients $e_{nm}$ have recently been derived up to the third order, $(\bar L/N^2)^3$, using computer algebra \cite{kuy95}. Another observable of interest in the study of superdeformed states is the $E2$ transitions. Assuming that the quadrupole transition operator is the same as in the Hamiltonian, i.e. $T(E2)=e{\bf Q}$ where $e$ is an effective boson charge, the $E2$ matrix elements are given by \begin{equation} \langle L \parallel T(E2) \parallel L-2 \rangle = e N \hat L \langle L0\, 20|L-2\ 0\rangle \bigl[m_1 + m_2 L(L-1)/N^2\bigr] \label{e2} \end{equation} where $\hat L = [2L+1]^{1/2}$ and the coefficients $m_n$ are given in Ref.~\cite{kuy95}. The first term in Eq. (\ref{e2}) gives the familiar rigid-rotor result. The second term is negative and is responsible for the boson cutoff effect in $E2$ transitions. Before applying the 1/$N$ expansion results, we compare them with those obtained from an exact diagonalization of the Hamiltonian. Diagonalization is carried out for $N=10$ which is the maximum possible boson number for this purpose. The quadrupole parameters $q_{22}, q_{24}, q_{44}$ are scaled down from their SU(3) values with $q=0.7$ which gives an adequate parametrization for the Hg-Pb region. Fig.~\ref{fig2}-a compares exact results for the dynamic moment of inertia ${\cal J}^{(2)}$ (circles) with three different $1/N$ calculations. The solid line shows the third order VAP results which is seen to follow the exact results very accurately. The second order VAP (dotted line) and the third order VBP results (dashed line) break down around spin $L\sim 2N$. Hence for description of high-spin states, the third order $1/N$ expansion formulas with VAP seem to be both necessary and sufficient. In Fig.~\ref{fig2}-b, the exact $E2$ transition matrix elements (circles) are compared with those obtained from Eq. (\ref{e2}) (line). The agreement is again very good up to very high-spins. Note that the accuracy of the $1/N$ expansion results improves with increasing $N$, hence in actual applications with $N\sim 30$, one would expect an even better agreement. The test case discussed here indicates that the extended formalism can be applied with confidence in the spin region $L=N$-$3N$ which covers the presently available data range for superdeformed bands. \section{APPLICATIONS TO SUPERDEFORMED BANDS} In this section, we apply the $1/N$ expansion formulas first in a systematic study of dynamic moment of inertia and $B(E2)$ values, and then to describe the experimental data on superdeformed bands. Since $\kappa$ is a scale parameter for energies, we need to study the effect of the three quadrupole parameters $q_{22}$, $q_{24}$, $q_{44}$. Fig.~\ref{fig3} shows the effect of variations in each $q_{jl}$ on dynamic moment of inertia while the other two are held constant at the SU(3) values. Here $q$ denotes the scaling parameter from the SU(3) values. Thus $q=1$, corresponds to the SU(3) limit which exhibits the rigid-rotor behaviour. To describe the variations in ${\cal J}^{(2)}$, one needs to break the SU(3) limit. From Fig.~\ref{fig3} it is seen that ${\cal J}^{(2)}$ is most sensitive to $q_{24}$ (note the different scales in the three figures). The other (diagonal) parameters have smaller and opposite effect on ${\cal J}^{(2)}$. Since the amount of data does not justify use of too many parameters, we prefer to scale all three with the same parameter $q$. The result of this simultaneous scaling is shown in Fig.~\ref{fig4}-a which is essentially the same as the one for $q_{24}$ in Fig.~\ref{fig3}. An interesting feature of these results is that the quadrupole Hamiltonian has the scope to describe both the increases and decreases in ${\cal J}^{(2)}$. For $q<1$, the $s-d$ coupling is relatively stronger than the $d-g$ coupling which results in loss of monopole pairing with increasing spin, and hence increase in ${\cal J}^{(2)}$. The opposite happens for $q>1$. Fig.~\ref{fig4}-b shows the effect the simultaneous variations in the quadrupole parameters on the $B(E2)$ values. The curving down of lines is due to boson cutoff which is most effective for smaller values of $q$. In the light of the above systematic studies, we have carried out fits to the available data on superdeformed bands in the Hg-Pb and Gd-Dy regions. The boson number is determined from microscopics, and $\kappa$ and $q$ are fitted to the data. The parameter values are given in the figure captions and the data are taken from the compilation in Ref. \cite{fir94}. Figs.~\ref{fig5} and \ref{fig6} compare the experimental dynamic moment of inertia (circles) with the calculated ones (lines) in Hg and Pb isotopes, respectively. In all cases ${\cal J}^{(2)}$ exhibits a smooth increase which is well reproduced by the calculations. The situation in the Gd-Dy region is not as favorable for our simple collective model as the other region, because there are definite signs indicating the importance of the single particle degree of freedom. For example, in $^{144-146}$Gd there are sudden jumps in ${\cal J}^{(2)}$ which are probably due to particle alignment effects \cite{lun94}. In $^{148-150}$Gd, ${\cal J}^{(2)}$ behaves reasonably smoothly so we have attempted to describe them (see Fig.~\ref{fig7}). The average behaviour of ${\cal J}^{(2)}$ in $^{148}$Gd is reproduced but the model fails in the case of $^{150}$Gd, underscoring the importance of single particle degree of freedom. For a better description of the data, one needs to incorporate particle alignment effects in the present formalism by including two-quasiparticle states in the basis \cite{iac91}. The dynamic moment of inertia of superdeformed bands in Dy isotopes exhibit an entirely different behaviour (Fig.~\ref{fig8}). They are very close to the rigid-rotor values, and hence the SU(3) limit as reflected in the values of $q\sim 1$. The $B(E2)$ values provide a complimentary observable to ${\cal J}^{(2)}$ which could be used as a further test of the model. In Fig.~\ref{fig9}, the available $B(E2)$ data in $^{192-194}$Hg (circles) are compared with the calculations. A reasonable description is obtained using boson effective charges $e=0.12-0.14~eb$ which are typical values used in the normal IBM calculations. Since the $B(E2)$ values are sensitive to the boson number (they vary as $N^2$), this provides a consistency check on the microscopically derived boson numbers. A further $N$ dependence is provided by the boson cutoff term in Eq. (\ref{e2}) which causes a drop in the calculated $B(E2)$ values at high-spins. Least-square fits to the data indeed indicate a drop in the $B(E2)$ values towards the high-spin end. However, the error bars are too large to reach an unambiguous conclusion whether this effect is genuine or not. \section{CONCLUSIONS} In this paper, we have reviewed a microscopic basis and a practical formulation of the IBM for application to superdeformed nuclei. The availability of analytical formulas owing to the 1/$N$ expansion technique means fast and efficient analysis of data. As first examples, we have considered the superdeformed bands in the Hg-Pb and Gd-Dy regions. A good description of data is obtained in the Hg-Pb region confirming the simple quadrupole nature of these superdeformed bands. In the Gd-Dy region, the dynamic moment of inertia exhibits large variations which can not be accommodated in a simple collective model. Such variations are due to the single particle degree of freedom and require extension of the model for a better description of the data. Finally, the formalism can be used in investigating some other interesting features of superdeformed nuclei such as identical bands and $C_4$ symmetry which will be pursued in future work. \section{ACKNOWLEDGEMENTS} This work is supported in parts by an exchange grant from the Australian Academy of Science/Japan Society for Promotion of Science and by the Australian Research Council, and in parts by Grant-in-Aid for Scientific Research on Priority Areas (No. 05243102) from the Ministry of Education, Science and Culture.
\section{#1}} \textwidth 159mm \textheight 230mm \def\bf Z {\bf Z} \def\tilde{\chi} {\tilde{\chi}} \def\sigma {\sigma} \def\bar{q} {\bar{q}} \def\epsilon {\epsilon} \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \def\begin{array}{\begin{array}} \def\end{array}{\end{array}} \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \def\begin{eqnarray}{\begin{eqnarray}} \def\end{eqnarray}{\end{eqnarray}} \def\rightarrow{\rightarrow} \def\longrightarrow{\longrightarrow} \def\hspace{0.1in}{\hspace{0.1in}} \def\hspace{0.1in}{\hspace{0.1in}} \def\hspace{0.2in}{\hspace{0.2in}} \def\hspace{0.3in}{\hspace{0.3in}} \def\hat{S}{\hat{S}} \def{\cal C}{{\cal C}} \def{\cal O}{{\cal O}} \def{\cal F}{{\cal F}} \def{\cal K}{{\cal K}} \def{\cal M}{{\cal M}} \def{\cal P}{{\cal P}} \def{\cal T}{{\cal T}} \def\Delta{\Delta} \defU_q({\cal G}){U_q({\cal G})} \begin{document} \oddsidemargin 5mm \renewcommand{\thefootnote}{\fnsymbol{footnote}} \newpage \begin{flushright} DAMTP-HEP-95/26 \end{flushright} \vspace{0.2cm} \begin{center} {\large {\bf Local Operators in Massive Quantum Field Theories} }\\ \vspace{0.3cm} {\bf Anni Koubek\footnote{Work supported by PPARC grant no. GR/J20661} }\\ \vspace{0.2cm} {\em Department of Applied Mathematics and Theoretical Physics, \\ CB3 9EW Cambridge, UK } \end{center} \subsection*{\large\bf 1 \phantom{11} Introduction} A fundamental problem in quantum theory is to establish a connection between its local description (quantum field theory) and measurable quantities (particle masses, scattering amplitudes). In elementary particle physics one relies mostly on approximation techniques due to the non-integrable structure of the interactions. In order to gain a deeper understanding of quantum theory in general these issues are examined for integrable systems, where one hopes to gain exact relations between the two descriptions. Though many integrable theories in four dimensions are known (for a review see e.g. \cite{fourd}) much more knowledge has been obtained so far for two dimensional theories, and it is thus this class of models where my investigations focus. In the particle picture the interaction is encoded into the scattering matrix. The asymptotic states are described by a linear superposition of free one-particle states $\vert Z_\epsilon (\beta)\rangle$, which are characterised by the particle species $\epsilon$ and their momentum, parametrised as $p^{(0)} =m \cosh \beta$, $p^{(1)} =m \sinh \beta$ ($m$ denotes the mass and $\beta$ the rapidity). They are related through the $S$-matrix as \begin{equation} \vert Z_{\epsilon_1}(\beta_1) \dots Z_{\epsilon_n}(\beta_n) \rangle_{in} = S_{\epsilon_1\dots\epsilon_n}^{\epsilon'_1\dots\epsilon'_m}(\beta_1,\dots,\beta_n \vert \beta_1'\dots\beta'_m ) \, \vert Z_{\epsilon'_1}(\beta'_1) \dots Z_{\epsilon'_m}(\beta'_m) \rangle_{out} \hspace{0.2in} .\end{equation} On the other hand the local description of a theory consists of the space of local operators ${\cal A} = \{ {\cal O}_i \} $ and the set of multi-point correlation functions of them, $$ \langle 0\vert {\cal O}_1(x_1 ) \dots {\cal O}_n(x_n ) \vert 0 \rangle \hspace{0.2in} .$$ The two description are linked through the Lehmann-Symanzik-Zimmermann reduction, since the particles can be obtained as asymptotic limits of the local fields. Another connection is given through the form factors. Consider an arbitrary two-point correlation function \[ G_{ij}(x)\,=\,<{\cal O}_i(x)\,{\cal O}_j(y)> \hspace{0.2in} , \] of hermitian operators. Inserting the Identity between the two operators and expanding it into the base of asymptotic states, it can be expressed as an infinite series over multi-particle intermediate states, \begin{equation} \langle {\cal O}_i(x)\,{\cal O}_j(y)\rangle\,=\label{correlation}\end{equation} $$ \sum_{n=0}^{\infty} \int \frac{d\beta_1\ldots d\beta_n}{ (2\pi)^n} <0|{\cal O}_i(x)|Z_{\epsilon_1}(\beta_1),\ldots,Z_{\epsilon_n}(\beta_n)>_{\rm in}{}^{\rm in} <Z_{\epsilon_1}(\beta_1),\ldots,Z_{\epsilon_n}(\beta_n)|{\cal O}_j(y)|0> $$ The matrix elements $$<0|{\cal O}_i(0)|Z_{\epsilon_1}(\beta_1),\ldots,Z_{\epsilon_n}(\beta_n)> = F_{\epsilon_1\dots\epsilon_n}(\beta_1,\dots,\beta_n)$$ are called form factors and in the following I will try to explain how they can be used in order to establish a link between the local description and the particle picture of a theory. \subsection*{\large\bf 2 \phantom{11} Form Factors and the Space of Local Operators} If one considers two dimensional integrable theories many simplifying properties occur which allow to calculate many dynamical quantities exactly. The most remarkable fact is the factorisation of the $S$-matrix, which determines a general scattering process as a product of two-particle scattering amplitudes. Further these two-particle $S$-matrices are pure phase-shifts, that is the incoming and outgoing momenta are the same. This simplification allows to calculate the $S$-matrix exactly (see {\em e.g.} P. Kulish's lectures in these proceedings). Also the form-factors can be determined exactly for integrable two dimensional systems. They obey a set of constraint equations, originating from fundamental principles of quantum theory, such as unitarity, analyticity, relativistic covariance and locality \cite{Karowski,nankai}. The important fact is that the $S$-matrix is the only dynamical information needed. In the following I will discuss just two examples of form-factor equations, in order to determine their overall structure, and also to explain the solution techniques. Since the theories here considered are defined in only one space dimension, a scattering process can be viewed as to interchange two particles on the real line, \begin{equation} Z_{\epsilon_1}(\beta_1) Z_{\epsilon_2}(\beta_2) = S_{\epsilon_1\epsilon_2}(\beta_1-\beta_2) Z_{\epsilon_2}(\beta_2) Z_{\epsilon_1}(\beta_1) \hspace{0.2in} .\end{equation} This exchange property will lead to a constraint equation for the form-factors \begin{eqnarray} & & F^{\cal O}_{\epsilon_1\dots\epsilon_i\epsilon_{i+1}\dots\epsilon_n} (\beta_1,\dots,\beta_i,\beta_{i+1},\dots\beta_n) =\nonumber\\ & &S_{\epsilon_i\epsilon_{i+1}}(\beta_i-\beta_{i+1})\, F^{\cal O}_{\epsilon_1\dots\epsilon_{i+1}\epsilon_{i}\dots\epsilon_n} (\beta_1,\dots,\beta_{i+1},\beta_{i},\dots\beta_n) \hspace{0.2in} . \label{wat1}\end{eqnarray} Another constraint equation derives from the bound state structure of the theory under consideration. If particles $Z_i$, $Z_j$ form a bound state $Z_k$, the corresponding two-particle scattering amplitude exhibits a pole at $\beta = iu_{ij}^{k}$ with the residue \begin{equation} -i \lim_{\beta \rightarrow iu_{ij}^{k}}(\beta-iu_{ij}^{k}) S_{ij}(\beta) =(\Gamma_{ij}^{k})^2 \hspace{0.2in} ; \label{residueS}\end{equation} $\Gamma_{ij}^{k}$ is the three--particle on--shell vertex. Corresponding to this bound state the form--factor exhibits a pole with the residue $$ -i \lim_{\beta'\rightarrow \beta}(\beta'-\beta) F^{\cal O}_{\epsilon_1\dots ij\dots\epsilon_n} (\beta_1,\dots,\beta'+i(\pi- u_{ik}^{j}),\beta-i(\pi- u_{jk}^ {i}),\dots,\beta_{n}) = $$ \begin{equation} =\Gamma_{ij}^{k} F^{\cal O}_{\epsilon_1\dots k\dots\epsilon_n} (\beta_1,\dots,\beta,\dots,\beta_{n} ) \hspace{0.2in} . \label{bounds}\end{equation} As mentioned before, (\ref{wat1}) and (\ref{bounds}) are only two examples of form-factor equations. Nevertheless they are two exponents of the two categories of the constraint equations: \begin{enumerate} \item Equations with fixed n ( e.g. (\ref{wat1})): they involve form factors with the same number of particles on both sides of the equation \item Recursive equations ( e.g.(\ref{bounds})): They link form factors with different particle numbers with each other - in the example above $n+1$ particle form factors to $n$ particle form factors. \end{enumerate} For theories with scalar particles there exists a well established solution method \cite{Karowski}. It consists of the ansatz \begin{equation} F_{\epsilon_1\dots\epsilon_n}(\beta_1,\dots,\beta_n) =Q_{\epsilon_1\dots\epsilon_n}(e^{\beta_1},\dots,e^{\beta_n}) \prod_{i<j}^n F_{\epsilon_i\epsilon_j} (\beta_i -\beta_j)\hspace{0.2in} . \end{equation} The two-particle form factors $F_{\epsilon_i\epsilon_j}$ can be calculated easily from the form factor equations. The product term satisfies all equations of the first type (with fixed particle number) and also is designed in order to have the correct pole structure of an $n$-particle form factor. Through this parametrisation the form-factor equations are reduced to recursive relations for the functions $Q_{\epsilon_1\dots\epsilon_n}(e^{\beta_1},\dots,e^{\beta_n})$. Further properties of these functions can be extracted from the form factor equations: they are homogeneous polynomials, symmetric in repeated indices with a total degree in its arguments fixed by relativistic covariance and the partial degree determined from the recursion relations. This information is sufficient for simple models to obtain explicit expressions for the form factors. In all cases though it is possible to determine the space of local operators by just considering these general properties of the functions $Q$ \cite{mywork}. Each linear independent solution of the form factor equations corresponds to an independent local operator. Therefore the space of local operators can be determined by counting the number of independent solutions of the form factor equations. This can be done due to the property of the recursion relations, that the dimension of the solution space at level $n$ is the sum of the dimension of the solution space at level $n-1$ and of the dimension of the kernel of the recursion relation at level $n$. Symbolically this can be written as $$dim(Q_n) = dim(Q_{n-1}) + dim ( {\cal K}_n)\hspace{0.2in} .$$ \begin{figure}[hbt] \begin{center} \begin{picture}(330,200)(-30,-30) \thicklines \put(0,0){\circle*{8}} \put(0,0){\line(1,0){200}} \put(0,0){\line(0,1){180}} \put(0,-15){\makebox(0,0){${\cal M}_{3}$}} \put(0,-30){\makebox(0,0){$h,\bar{h} =\{ 0,\frac 1{16},\frac 12\}$}} \put(200,-15){\makebox(0,0){$H=0$, $T>T_c$}} \put(100,35){Free fermion theory $S=1$} \put(100,15){Local operators: $1 \, (0,0)$, $\psi \, (0, \frac 12)$, $\bar{\psi} \, ( \frac 12,0)$, $\epsilon \, (\frac 12, \frac 12)$} \put(-25,170){\makebox(0,0){$H \neq 0$}} \put(-25,150){\makebox(0,0){ $T=T_c$}} \put(10,160){$E_8$-scattering theory} \put(10,140){Local operators: $1 \, (0,0)$, $\sigma \, (\frac 1{16}, \frac 1{16})$, $\epsilon \, (\frac 12 ,\frac 12)$} \end{picture} \caption{The critical Ising model and its integrable perturbations} \label{fig2} \end{center} \end{figure} An interesting application of this counting method is in perturbed conformal field theories. As an example consider the Ising model. Its point of second order phase transition is described by a conformal field theory ( the minimal model ${\cal M}_{3}$ ) and therefore the space of local operators at the critical point is determined by Virasoro irreducible representations. The critical point admits two relevant perturbations which are integrable. The perturbation with the conformal operator with conformal weight $h=\frac 12$ drives the model into the regime $T>T_c$ and $H=0$. It is described by a free fermion theory. The other perturbation with the operator $h=\frac 1{16}$ corresponds to the Ising model with $H \neq 0$ and $T=T_c$ and is described by a scattering theory containing 8 scalar massive particles \cite{Zam}. Virasoro symmetry is obviously broken by both perturbations and it is therefore an interesting problem to determine the space of operators for these theories. The situation is summarised in figure \ref{fig2}. For both perturbations it is possible to determine the space of local operators. It is given in terms of characters of the minimal model $M_{3}$. This is a quite remarkable fact, since conformal symmetry is explicitly broken by the perturbation. Further, since the thermal perturbation is described by a free fermion theory, the local operators are just the fermions $(\psi, \bar{\psi})$ the identity operator ($1$) and the energy density $\epsilon$. Also the spin operator ($\sigma$) and the disorder field ($\mu$) can be analysed by the counting method, but they are semi-local operators with respect to the asymptotic states. The magnetic perturbation breaks the $Z_2$ symmetry of the model and only scalar operators appear in this perturbation, namely the identity ($1$), the energy density ($\epsilon$) and the spin operator ($\sigma$). A general feature of the counting method is that the space of operators is determined by fermionic sum expressions. Such expressions also appear in the analysis of corner transfer matrices \cite{kedem} and spinon conformal field theories \cite{spinon}. It would be interesting to establish a more direct connection between these methods and the form factor approach. Finally note that the above example only constitutes a simple application of the counting method. It can be generalised to many other systems, including models with a massless spectrum and/or bounderies.
\section{Introduction} Quantum localization is a generalization Duistermaat-Heckman theorem \cite{dh} to infinite dimensions. This theorem states that if the Hamiltonian $H$ generates a global circle, or, more generally a torus action in the phase space $\Gamma$ then the canonical partition function is given exactly by the saddle-point approximation around the critical points of $H$. Extensions to calculation of quantum mechanical partition functions using phase space path integrals have been represented in {e.g.} \cite{us}. We shall first consider basic ideas of localization. Then we shall carefully regularize the pertinent functional determinants arising from the path integrals. There is an ambiguity in choosing the regularization scheme because of the spectral asymmetry of first order differential operators. Therefore, the result depends on the regularization as in the case of quantum mechanical anomalies. Finally, we are going to apply our localization to the quantization of the simple harmonic oscillator and to the evaluation of the Weyl character of spin. We shall notice that different regularizations give different energy spectra for the harmonic oscillator. We also show that the continuum coherent state path integral yields directly the correct character for spin if we choose an appropriate regularization. In particular, we will consider the relation of character formulae to the Borel-Weil theory which constructs the irreducible representations of a Lie group as holomorphic functions. Using this theory we relate the character formulae to the equivariant index of the Dolbeault complex. The result is that the path integral yields directly the correct character without an explicit Weyl shift of the highest weight. \\ \section{Localization of Phase Space Path Integrals} We are interested in exact evaluation of phase space path integrals (partition functions) of the form \begin{equation}\label{Zcanonical} Z(T) = \int_{L \Gamma} {{\cal D}}x \ {\rm Pf \ } \Vert \omega_{ab}(x) \Vert \exp \left( \ii \int_0^T \dd t \left[ \vartheta_a {\dot{x}}^a - H(x) \right] \right) \;. \end{equation} where $\{x^a\} $ are local coordinates in $\Gamma$, ${\rm Pf \ } \Vert \omega_{ab} \Vert$ is the Liouville measure factor, $\vartheta_a$ the symplectic potential and $\omega_{ab} = {\partial}_a \vartheta_b - {\partial}_b \vartheta_a$. The integration is performed over the loop space $L\Gamma$ consisting of the phase space loops. The integrability condition \cite{woodhouse} requires that $$\int_{\Sigma} \omega = 2 \pi n$$ for any 2-cycle $\Sigma$ in $\Gamma$ so that the path integral is single valued. We introduce anticommuting variables $\psi^a$ to write ${\rm Pf \ } \Vert \omega_{ab} \Vert $ as a path integral \begin{equation}\label{ZSUSY} Z(T) = \int {{\cal D}}x {{\cal D}}\psi \ \exp \left( \ii \int_0^T \dd t \left[\vartheta_a {\dot{x}}^a - H(x) + \half \psi^a \omega_{ab} \psi^b \right] \right) \; . \end{equation} The boundary conditions are periodic also for the fermions, since they are a realization of the differentials of the bosonic coordinates. We interpret the path integral (\ref{ZSUSY}) in terms of equivariant cohomology in $L\Gamma$. From the bosonic part of the action we get a Hamiltonian vector field in $L\Gamma$ \begin{equation} {\chi}_S^a = \dot{x}^a - \omega^{ab} {\partial}_b H $$ whose zeroes define the Hamilton's equations. The equivariant exterior derivative in $L\Gamma$ is \begin{equation} \dd_S = \dd + \iota_S \;, $$ where $\iota_S$ denotes the contraction along the vector field $\chi_S$. The square of $\dd_S$ is the loop space Lie derivative $$% {\cal L}_S = \dd \iota_S + \iota_S \dd \sim {\dd \over \dd t} - {\cal L}_H \;. $$ The action $S_{\rm B} + S_{\rm F}$ is supersymmetric under the infinitesimal loop space supersymmetry transformations that are parametrized by a gauge fermion $\delta \Psi$: \begin{eqnarray}\label{SUSY} x^a \rightarrow x^a + \delta \Psi \dd_S x^a & = & x^a + \delta \Psi \psi^a , \cr \psi^a \rightarrow \psi^a + \delta \Psi \dd_S \psi^a & = & \psi^a + \delta \Psi \chi_S^a \;. \end{eqnarray} This implies that that the action is equivariantly closed: $$\dd_S (S_{\rm B} + S_{\rm F}) =0 \;.$$ By an analogue of Fradkin-Vilkovisky theorem \cite{bfv} one can show that the path integral remains intact if we modify the action by $S \rightarrow S+\dd_S \Psi$ where $\Psi$ satisfies the Lie derivative condition \begin{equation}\label{EquivCohomology} \dd_S^2 \Psi = {{\cal L}}_S \Psi =0 \;. \end{equation} In the limit $\lambda \rightarrow 0$ the path integral \begin{equation}\label{Z_lambda} Z_{\lambda} (T) = {\cal} \int {{\cal D}}x^a {{\cal D}}\psi^a \ \exp \left( \ii \int_0^T \dd t \left[\vartheta_a {\dot{x}}^a - H(x^a) + \half \psi^a \omega_{ab} \psi^b + \lambda \dd_S \Psi \right] \right) \end{equation} reduces to (\ref{ZSUSY}) and $\lambda \rightarrow \infty$ gives localization. To construct a gauge fermion $\Psi$ we need a metric $g$ in the phase space. The loop space Lie derivative condition (\ref{EquivCohomology}) is satisfied if the metric $g$ in $\Gamma$ is invariant under the Hamiltonian action of $H$ \begin{equation}\label{InvariantMetric} {{\cal L}}_H g = 0 \;, \end{equation} which means that $\chi_H$ is a Killing vector field. This is a very restrictive condition for the Hamiltonian: it must generate a global U(1)-action in $\Gamma$. We can choose any metric which satisfies the condition (\ref{InvariantMetric}) and average it over the group action. We will consider the following selections for the gauge fermion: \begin{equation} \Psi_1 = {1 \over 2} g_{ab} \dot{x}^a \psi^b $$ gives localization to the constant modes which are points of the manifold, \begin{equation} \Psi_2 = {1 \over 2} g_{ab} \chi_H^a \psi^b $$ to the zeroes of $\chi_H$ which we assume to be nondegenerate and isolated and \begin{equation} \Psi_3 = {1 \over 2} g_{ab} \chi_S^a \psi^b $$ to the classical trajectories. For simplicity we use subscripts 1,2,3 in the actions and partition functions corresponding to the gauge fermions $\Psi_{1,2,3}$. The actions become \begin{eqnarray} S_1 & = & \int_0^T \dd t \ \left[ \left( \vartheta_a - {\lambda \over 2} g_{ab} {\chi}_H^b \right) \dot{x}^a -H + {\lambda \over 2} g_{ab} \dot{x}^a \dot{x}^b + {\lambda \over 2} \psi^a \left( g_{ab} {{\partial}}_t + \dot{x}^c g_{bd} {\Gamma}_{ac}^d \right) \psi^b + {1 \over 2} \psi^a \omega_{ab} \psi^b \right] \;, \cr S_2 & = & \int_0^T \dd t \ \left[ \vartheta_a \dot{x}^a - H + {\lambda \over 2} g_{ab} {\chi}_H^a \left( \dot{x}^b - \chi_H^b \right) + {\lambda \over 2} \psi^a {\partial}_{a }\left( g_{cb} \chi_H^c \right) \psi^b + {1 \over 2} \psi^a \omega_{ab} \psi^b \right] \;, \cr S_3 &= & \int_0^T \dd t \left[\vartheta_a \dot{x}^a - H +{\lambda \over 2} g_{ab} \chi_S^a \chi_S^b + {\lambda \over 2} \psi^a {\partial}_a (g_{cb} \chi_S^c) \psi^b + {1 \over 2} \psi^a \omega_{ab} \psi^b \right] \;. \end{eqnarray} To take the limit $\lambda \rightarrow \infty$ in path integrals we make the decomposition to constant modes $x_0^a, \psi_0^a$ and to non-constant modes $x^a_t , \psi^a_t$ and scale the non-constant modes by $1/ \sqrt{\lambda} $ \begin{eqnarray}\label{scaling} x^a(t) & = & x_0^a + {1 \over {\sqrt{\lambda}}}x^a_t , \cr \psi^a(t) & = &\psi^a_0 + {1 \over {\sqrt{\lambda}}} \psi^a_t . \end{eqnarray} The Jacobi determinant is unity. An expansion to a quadratic order around the constant modes and the limit $\lambda \rightarrow \infty$ gives a Gaussian path integral \begin{equation}\label{xdot-Gaussian} Z_1 = \int \dd x_0^a \dd \psi_0^a \ \exp \left[ -\ii T \left( H- \half \psi^a_0 \omega_{ab} \psi^b_0 \right) \right] Z_{\rm fl,1} (T) \end{equation} where the fluctuation path integral $Z_{\rm fl}(T) $ is a product of fermionic and bosonic parts: \begin{eqnarray}\label{x-Fluctutation} Z_{\rm F ,1} &=& \int \prod_t \dd \psi^a_t \ \exp \left\{-{ \ii \over 2}\int_0^T \dd t \ \psi^a_t g_{ab} {{\partial}}_t \psi^b_t \right\} \; , \cr Z_{\rm B,1} & = & \int \prod_t \dd x^a_t \ \exp \left\{ { \ii \over 2} \int_0^T \dd t \ x^a_t \left[ {\cal R}_{ab} {{\partial}}_t -g_{ab} {{\partial}}_t^2 \right] x^b_t \right\} \; . \end{eqnarray} Here $${\cal R}_{ab} = R_{ab} + \tilde{\Omega}_{ab}$$ is the equivariant curvature with $R_{ab} $ the Riemannian curvature 2-form and $$ \tilde{{\Omega}}_{ab} = \half \left[ {\nabla}_{b}(g_{ac} \chi_H^c) - {\nabla}_{a}(g_{bc} \chi_H^c) \right] $$ the momentum map \cite{arnold} corresponding to ${\chi}_H$, $\nabla$ being the covariant derivative. $H$, $\omega$, $g$ and ${{\cal R}}$ are evaluated at the constant modes. The path integral $Z_2$ is given by a sum over the critical points $\{x_i \}$ of the Hamiltonian: \begin{equation}\label{FPICritical} Z_2 = \sum_{ x_i } \ { {\exp[-\ii TH] } \over { {\rm Pf \ } \Vert {\partial}_a \chi_H^b \Vert } } Z_{\rm fl,2}(T) \;. \end{equation} $Z_{\rm fl,2}$ is also a product of fermionic and bosonic parts: \begin{eqnarray}\label{chi-Fluctutation} Z_{\rm F,2}(T) & = & \int \prod_t \dd \psi^a_t \ \exp \left\{ {\ii \over 2} \int_0^T \dd t \ \psi^a_t {\partial}_a (g_{bc} \chi_H^c) \psi^b_t \right\} \;, \cr Z_{\rm B,2}(T) & =& \int \prod_t \dd x ^a_t \ \exp \left\{ {\ii \over 2} \int_0^T \dd t \ x^a_t {\partial}_a (g_{bc} \chi_H^c) (\delta^b_d {\partial}_t - {\partial}_d \chi^b_H ) x^d_t \right\} \;. \end{eqnarray} Here $g$ and $\chi_H$ are again evaluated at the constant modes. Finally, the path integral $Z_3$ reduces to a sum over the $T$-periodic classical trajectories \begin{equation} \label{Classical} Z_3 = \sum_{x_{\rm {cl}}} { 1 \over {\rm Pf \ } \Vert \delta_b^a {\partial}_t - {\partial}_b \chi_H^a \Vert } \exp[\ii S_{{\rm cl }}] \;. \end{equation} In practice, it is usually a highly non-trivial problem to find the $T$-periodic classical trajectories of a dynamical system \cite{ArnoldConj}. \\ \section{Regularization of Fluctuation Path Integrals} In the following all the path integrals and determinants are evaluated over periodic configurations for both the bosonic and fermionic degrees of freedom. The primes will denote that we exclude the constant modes. In real polarization the fluctuation parts in $Z_{1,2}$ become \begin{eqnarray}\label{xdot-Z_fl} Z_{\rm fl ,1} & = & {1 \over \sqrt{{\rm Det\ } ' \Vert \delta_b^a {\partial}_t - {{\cal R}}_b^a \Vert }} \;, \cr Z_{\rm fl,2} &= & {1 \over \sqrt{{\rm Det\ } ' \Vert \delta_b^a {\partial}_t - {\partial}_b \chi_H^a \Vert}} \;. \end{eqnarray} It is quite important to notice that in the reduced determinants one index is covariant and another contravariant. In K\"ahler polarization the fluctuations parts are, using the additional symmetries of the metric and the Riemann curvature tensor \cite{nakahara}, \begin{eqnarray} Z_{\rm fl,1} & = & {1 \over {\rm Det\ }' \Vert \delta_a^b {\partial}_t -{\cal R}_a^b \Vert }\;, \cr Z_{\rm fl,2} & = & {1 \over {\rm Det\ } ' \Vert \delta_a^b {\partial}_t - {\partial}_a \chi_H^b \Vert } \;. \end{eqnarray} These determinants are taken over the holomorphic indices. By this we mean the following: The relevant matrices can be block diagonalized \begin{equation} A = {\rm diag} \ (A_1 ,A_2 ,... ,A_N) $$ with blocks \begin{equation} A_k = \pmatrix{ a^+_k & 0 \cr 0 & a^-_k \cr } \equiv \pmatrix{ a_k & 0 \cr 0 & - a_k \cr } \;. $$ The symbols $a^+_k$ and $a^-_k$ denote the holomorphic and antiholomorphic eigenvalues of $A$, and we consider only the eigenvalues corresponding to the holomorphic indices to the determinant. We have to choose a regularization scheme for the determinants. A standard method is to apply $\zeta$- and $\eta$-functions. The $\zeta$-function regularization does not directly apply to first-order operators because they have an infinite number of negative eigenvalues. To take them into account we define the $\eta$-function for the first-order operator $B$ by \begin{equation} \eta_B (s) = \sum_{b_{n} \neq 0} {{\rm sign\ } } (b_{n}) \abs{b_{n}}^{-s} + {\rm dim \ Ker \ } B = {1 \over {\Gamma\left({s+1} \over 2 \right)}} \int_0^{\infty} \dd t \ t^{(s-1)/2} \ {\rm Tr\ } [B \exp(-t B^2)] \;. $$ Analytical continuation to $s=0$ gives the Atiyah-Patodi-Singer $\eta$-invariant \cite{egh} of $B$ that measures the spectral asymmetry of $B$ and specifies the phase of ${\rm Det\ }(B)$. The absolute value $\abs{{\rm Det\ } (B) }$ is regularized using the formula $$% \abs{{\rm Det\ } (B)} = +\sqrt{{\rm Det\ } (B^2)} = + \exp\left[ - \half \zeta_{B^2} ' (0)\right] \;. $$ In real polarization we have to evaluate the square root of a determinant of the antisymmetric operator $B = {\partial}_t - A $ where $A$ is an antisymmetric matrix. In our case $A$ is $\Vert {\cal R}_a^b \Vert$ or $\Vert {\partial}_a \chi^b_H \Vert$. By determining the spectrum of $B$ and applying $\zeta$-function regularization we get, up to an inessential numerical normalization, the result \begin{equation}\label{Agenus} {1 \over {\sqrt{{\rm Det\ }' \left( {\partial}_t - A \right) }} } = \prod_{n= 1}^N { \abs{a_n /2 \over {\sin (a_n T/2 )}}}= {1 \over T^N} \hat{A}(TA) \;, \end{equation} where we have defined the function of the matrix $X$ $$\hat{A}(X) = \prod_n {x_n/2 \over \sin(x_n/2)} \;$$ where $x_n$ are the skew-eigenvalues of $X$. The result is non-negative since the negative and positive skew-eigenvalues appear in pairs. Therefore there is no ambiguity with the spectral asymmetry. Now we consider the determinants in K\"ahler polarization. It is sufficient to consider the determinant of a block. Earlier we noticed that the fluctuation path integrals reduce to the determinant of the operator $B = \ii {\partial}_t -a$. The functional Pfaffian in (\ref{Classical}) is also similar to this determinant. To regularize $${\rm Det\ }'(B) = \prod_{n \neq 0} \left( {2\pi n \over T} - a \right)$$ properly we have to take into account that $B$ has and infinite number of negative eigenvalues. Thus there is a problem with the spectral asymmetry. Therefore, we have to choose a regularization prescription which has a relation to quantum mechanical anomalies. In the regularization of the determinants it is not possible to maintain all the symmetries that are present in the classical theory. For example, Elitzur et al. \cite{elirab} considered the corresponding fermionic problem with antiperiodic boundary conditions. They evaluated the quantum mechanical partition function for a Dirac fermion in an external gauge field $A(t)$ in $0+1$-dimensions \begin{equation} Z(T) = \int {\cal D} \bar{\psi} {\cal D} \psi \exp\left[ \ii \int_0^T \dd t \ \bar{\psi} \left ( \ii {\partial}_t -a \right) \psi \right] = {\rm Det\ } ( \ii {\partial}_t -a ) \;. \end{equation} where, because of the gauge invariance of the action only the constant mode $a$ of $A(t)$ contributes. The classical action has both the invariance under large gauge transformations \begin{eqnarray} a &\rightarrow& a + n 2 \pi /T \cr \psi &\rightarrow& \psi \end{eqnarray} and the charge conjugation invariance \begin{eqnarray} a &\leftrightarrow& -a \cr \psi &\leftrightarrow &-\bar{\psi} \;. \end{eqnarray} However, when regularizing the determinant one has to choose which symmetry one wants to maintain, which leads to a global anomaly. Here we have an analogous situation. It is not {\em a priori} clear what the result of the regularization should be, and there is a genuine ambiguity. Since the zeroes of the determinant are at $aT = 2\pi n $, the determinant must be proportional to $$ {\sin( a T/ 2 ) \over a/2} \;. $$ The proportionality factor can be any function without zeroes that is, the exponent function. The determinant is therefore, up to an irrelevant constant,$${\rm Det\ }(\ii {\partial}_t -a ) = { \sin( a T/ 2) \over a/2} \exp\left(\ii \phi aT \right) $$ with a phase $\phi$ whose natural values turn out to be $0$ and $\pm 1/2$ since they yield the (anti)symmetries of the product under $a \leftrightarrow -a $ and $a \rightarrow a + 2\pi n /T$. However, there is a minor subtlety: in our localization formulae the zero modes are absent and this destroys these symmetries. Nevertheless, we may still consider the residual symmetries. The choice $\phi=0$ corresponds to neglecting the spectral asymmetry and choosing the (anti)symmetry $a \rightarrow -a$ to be unbroken. In this regularization scheme the inverse determinant is simply $$% { 1 \over {\rm Det\ }'( {\partial}_t - A) } = {1 \over T^N} \hat{A}(T A) $$ This is the result that usually appears in literature. However, there is another possibility. The values $\phi = \pm 1/2$ correspond to maintaining the symmetry $a \rightarrow a + 2\pi n /T$ and taking into account the spectral asymmetry by the Atiyah-Patodi-Singer $\eta$-invariant. This yields \begin{eqnarray}\label{ComplexDet} {1 \over {{{\rm Det\ } '}({\partial}_t - A)}} &=& \prod_{n= 1}^N {{a_n/2} \over {\sin \left( a_n T / 2 \right)}} \exp \left( \ii a_n T/2 \right) = {1 \over T^N} {{\rm Td } }(T A) \end{eqnarray} where we have defined the following function of the matrix $X$ $$ {\rm Td } (X) = \prod_n { x_n /2 \over \sin(x_n /2) } \e^{\ii x_n /2} \;. $$ We take only the eigenvalues corresponding to the holomorphic indices to the determinant. Let us now write down the resulting localization formulae. The localization to constant modes yields the expression \begin{equation}\label{ConstantModes} Z_1 (T) = {1 \over T^N} \int \dd x_0^a \dd \psi_0^a \ {{\rm Ch } }\left[ -\ii T(H- \omega) \right] \cases{ \hat{A}({T {\cal R} }) \cr {\rm Td } ({T {\cal R} }) } \;. \end{equation} We have defined equivariant generalizations \cite{bgv} of the conventional characteristic classes known as equivariant $\hat{A}$- and Todd-genus, and identified the exponential with the equivariant Chern class. When $H=0$ they reduce to the conventional characteristic classes and the result is a topological invariant. The localization to the critical points $\{ x_i\}$ of the Hamiltonian gives the result \begin{eqnarray}\label{CriticalPoints} Z_2 (T) = {1 \over T^N} \sum_{x_i} {\exp(-\ii TH ) \over {\rm Pf \ }({\partial} \chi_H^{~} )} \cases{ \hat{A}(T {\partial} \chi_H^{~} ) \cr {\rm Td } (T {\partial} \chi_H^{~} ) } \;. \end{eqnarray} We must use local coordinates in the evaluation of the determinants when localizing to the critical points of the Hamiltonian. Finally, the localization to $T$-periodic classical trajectories yields \begin{eqnarray}\label{WKB} Z_3 (T) = {1 \over T^N} \sum_{x_{\rm cl}} {\exp(\ii S_{\rm cl} ) \cases{ \hat{A}(T {\partial} \chi_H^{~} ) \cr {\rm Td } (T {\partial} \chi_H^{~} ) }} \;. \end{eqnarray} \section{ Harmonic Oscillator} Now we show that the localization formulae yield the correct partition function for the harmonic oscillator in a flat phase space. The path integral for it is Gaussian and in principle there is no reason to apply localization to it. However, it is reasonable to check by some simple examples that our assumptions and derivations are valid. In particular, we will show that the choice of the metric in the phase space is not relevant, contrary to claims in literature \cite{dlr}. It is also illustrative to consider the significance of the regularization schemes we have used. In real polarization the Hamiltonian is $H= \half \left( p^2 +q^2 \right)$ and the symplectic 2-form is $\dd q \wedge \dd p . $ The coherent state representation (K\"ahler polarization) requires some further investigation, since we have to fix an operator ordering prescription. In terms of creation and annihilation operators the normal and symmetric ordered Hamiltonians are, respectively, \begin{eqnarray}\label{Ordering} H_{\rm n} & =& : {1 \over 2} \left( a^{\dagger}a + a a^{\dagger} \right) : = a^{\dagger}a \;, \cr H_{\rm s} &=& a^{\dagger} a + \half \; . \end{eqnarray} The symmetric ordered Hamiltonian has an explicit zero point energy $E_0 = 1/2.$ To apply the localization formulae we must choose a metric in the phase space and calculate the equivariant curvature and the derivatives of the Hamiltonian vector field. If the Lie-derivative condition ${\cal L}_H g = 0$ is satisfied we can start from an any smooth metric in the phase space and average it. So we may choose a constant metric $$g = \pmatrix{ 1 & 0 \cr 0 & 1 \cr }.$$ The non-zero components of the equivariant curvature are \begin{equation} {\cal R}^p_q = - {\cal R}^q_p = 1 \;. $$ The localization formula (\ref{ConstantModes}) yields the result \b Z(T) & = & \int \dd p \ \dd q \ \dd \psi^q \ \dd \psi^p \ \exp \left[ -\ii T \left( {1 \over 2} p^2 + \half q^2 - \psi^p \psi^q \right) \right ] {1 \over 2\sin ( T/2)} \cr & \sim & {1 \over 2 \sin( T/2) } = \sum_{n=0}^{\infty} \exp [\ii (n + 1/2 )T ] \end{eqnarray} which is the correct partition function with the zero-point energy $E_0 = 1 /2 $. Let us now digress slightly to discuss the result. In \cite{dlr} Dykstra, Lykken and Reiten analyzed this problem and they noticed a dependence on the metric. What they did not notice was that the index structure of the equivariant curvature is ${\cal R}_a^b$ and therefore it is invariant under global scalings of the metric. Furthermore, they used a metric which in polar coordinates near the origin behaves like $$ \dd s^2 = \dd r^2 + c r^2 \dd \phi^2 \;. $$ This is a metric on a cone, not on a plane when $c \neq 1$ and is not smooth, nor even continuous at the origin. Therefore it is not surprising that their energy levels depend on the parameter $c$ which represents the tip angle of the cone. From this we indeed see that we cannot choose an arbitrary invariant metric, since it has to respect the topology of the phase space. The localization to the critical points of the Hamiltonian (\ref{CriticalPoints}) yields also the correct result. The only zero of $\chi_H^{~} $ is the origin of the phase space, which gives \begin{equation} Z(T) = {1 \over {\rm Pf \ } \pmatrix{ 0 & 1 \cr -1 & 0 \cr }} { {1 /2} \over {\sin ( T/2)}} = {1 \over { 2 \sin(T/2)} } \;. $$ If we want to apply the localization the classical trajectories we must classify all the $T$-periodic classical trajectories. If $T \neq 2 \pi n$ the problem reduces to the localization to the critical points of the Hamiltonian. However, if $T = 2 \pi n $ the zeroes of $\chi_S$ are not isolated and we have to use a degenerate version of the localization formula to the classical trajectories \cite{Palo} We now consider the harmonic oscillator in the K\"ahler polarization. We will only discuss the localization formulae to constant modes. The reasoning is similar with other formulae. There are four cases to consider: the localizations with the $\hat{A}$-genus and Todd-genus using two different orderings. We will only list the spectra we obtain. The use of $\hat{A}$-genus yields the spectra $E_n = n+1/2$ (normal ordering) and $E_n = n+1$ (symmetric ordering). The Todd-genus gives the results $E_n = n$ (normal ordering) and $E_n = n+1/2$ (symmetric ordering). The first and fourth results have the correct zero-point energy. From this example we see that to get correct results from the path integral we do need some additional information other than the classical action and boundary conditions: we must choose a regularization scheme that gives physically correct results. \section{ Character for SU(2)} We shall now use our localization formulae to derive the Kirillov and Weyl character formulae for Lie groups \cite{bgv}. The character formula for SU(2) has been widely discussed in literature \cite{afs, stone, nielsen}. However, there has been some controversy about the Weyl shift problem: the path integral usually gives almost the correct character up to the substitution $j \rightarrow j+1/2$. We show that the coherent state path integral and the localization formulae with the Todd-genus directly yield the correct character. In this calculation we use the continuum version of the coherent state path integral and show that this also yields the correct result, contrary to discussions in literature \cite{funahashi}. To motivate the use of Todd-genus we relate the character of a simple Lie group $G$ in the highest weight representation $\lambda$ to the index of the twisted Dolbeault complex on the coadjoint orbit $O_f$ \cite{perelomov} of the group. The Borel-Weil theory \cite{ stone, alsiwi} constructs the irreducible representations of $G$ as holomorphic sections of a line bundle $L$ that is associated to a principal bundle $ G \rightarrow G/T_G \sim O_f$ where $T_G$ is the Cartan torus of $G$ \cite{alsiwi}. The holomorphic sections of this line bundle (coherent states) form the basis for the irreducible representation. The connection 1-form on $L$ is the symplectic potential $$% \vartheta = {{\partial}{F} \over {\partial} z^k } \dd z^k - {{\partial}{F} \over {\partial} \bar{z}^k} \dd \bar{z}^k \; $$ where $F$ is the K\"ahler potential on $O_f$. It can be shown that the twisted Dolbeault operator $\bar{{\partial}}_L = \bar{{\partial}} + \vartheta_{\bar{z}}$ annihilates the normalized coherent states $\ket{z}$ and therefore $\ket{z} \in H^{0,0} (O_f , L)$. If we can prove that all the other cohomology groups are trivial, {e.g.} by Lichnerowicz vanishing theorem \cite{bgv}, we conclude that the dimension of the highest weight representation $R_{\lambda}$ is $\dim H^{0,0} (O_f, L)$. Consequently, this is equal to the index of the twisted Dolbeault complex. The Riemann-Roch-Hirzebruch index theorem relates this analytical index to the topological invariant \begin{equation}\label{TopDolbeault} {{\rm ind\ }} \ \bar{{\partial}}_L = {\dim}\ R_{\lambda} = \int_{O_f} {\rm Td}(O_f) \wedge {{\rm Ch } }(L) \;. \end{equation} Indeed, we notice that the localization formula (\ref{ConstantModes}) with $H=0$ represents this index provided we use the Todd-class. For SU(2) we obtain the known result for the dimension of the spin-$j$- representation $$ \dim R_j = {\rm ind\ } \ \bar{{\partial}}_L = 2j +1 \;, $$ This is the correct result without the explicit Weyl shift by the Weyl vector $\rho = 1/2$. We shall now use an equivariant version of the index theorem to derive the character formulae. The character of an element in the Cartan subalgebra is the partition function for the Hamiltonian $H$ that represents it on $O_f$: \begin{equation}\label{character} \chi (\beta) = {\rm Str \ } \exp[- \ii T H ] \;. \end{equation} To make a relation to the Dolbeault index we write this as an equivariant index (character index, G-index, Lefschetz number) \cite{bgv}. One can show that the Laplacians $\bar{{\partial}}_L^{\dagger} \bar{{\partial}}_L$ and $\bar{{\partial}}_L \bar{{\partial}}_L^{\dagger}$ have equal non-zero eigenvalues. If all other comohomology classes except $H^{0,0}$ are trivial, as we presume, $\bar{{\partial}}_L^{\dagger} $ does not have zero modes. Consequently, we can write the trace as an equivariant index: \begin{eqnarray} {\rm ind\ }_H (\bar{{\partial}}_L, T) &\equiv& \lim_{\beta \rightarrow \infty} {\rm Tr\ } \e^{- \ii T H} (\e^{- \beta \bar{{\partial}}_L^{\dagger} \bar{{\partial}}_L} - \e^{- \beta \bar{{\partial}}_L \bar{{\partial}}_L^{\dagger}} )\cr & =& \lim_{\beta \rightarrow \infty} {\rm Str \ } \exp[- \ii T H] \exp \left[- \beta \pmatrix{ \bar{{\partial}}_L^{\dagger} \bar{{\partial}}_L & 0 \cr 0 & \bar{{\partial}}_L \bar{{\partial}}_L^{\dagger} } \right] \end{eqnarray} Only the zero modes contribute to the trace. The expression is also independent of $\beta$. Therefore, in the limit $\beta \rightarrow 0$, all we are left with are the zero modes of $\bar{{\partial}}_L$. Consequently, the equivariant index is equal to the character $$ {\rm ind\ }_H (\bar{{\partial}}_L, T) = {\rm Str \ } \exp[- \ii T H] $$ Thus, the character is the equivariant index of the twisted Dolbeault complex and therefore we choose the localization with the Todd-class. To derive the character formulae we apply standard methods to write ${\rm Str \ } \exp[ - \ii T H]$ as a coherent state path integral of the form (\ref{ZSUSY}). Since we can choose an invariant metric on a coadjoint orbit \cite{perelomov} we can localize the path integral to classical trajectories, to constant modes or to critical points of the Hamiltonian. The two latter cases yield the Kirillov character formula \cite{kirillov} ($2N$ is the dimension of the orbit) \begin{equation}\label{Kirillov} \chi (T) = {1 \over T^N} \int_{O_f} {{\rm Ch } }\left[ -\ii T (H - \omega) \right] {\rm Td } ({ T {\cal R}^+}) \;, \end{equation} and the Weyl character formula \begin{equation}\label{Weyl} \chi (T) = {1 \over T^N} \sum_{z_i} {\exp(-\ii T H ) \over \det^+ ( {\partial} \chi_H^{~} ) } {\rm Td } (T {\partial} \chi_H^{~} ) \;, \end{equation} respectively. In (\ref{Weyl}) we have identified the Pfaffian in the real polarization with the determinant over the holomorphic eigenvalues of $\Vert {\partial}_a \chi^b_H \Vert $ and the summation is over the critical points of the Hamiltonian or equivalently the Weyl group. As the only example we evaluate the character for SU(2). We write the character as a coherent state path integral over the coadjoint orbit $\rm{SU(2)/U(1)} \sim S^2$. We choose complex coordinates by introducing the stereographic projection from the south pole. The K\"ahler potential on the orbit with radius $j$ is $F = j \log(1 +z \bar{z})$ from which we obtain the metric and the symplectic 1- and 2-forms in the standard fashion. The integrability condition requires $j$ to be a multiple of $1/2$: this is the topological quantization of spin. The canonical realization for $H = J_3$ is $$ J_3 = -j {{1 -z\bar{z}} \over {1+z\bar{z}}} \; $$ and the path integral for the character becomes (\ref{ZSUSY}) \begin{equation}\label{SU2Character} \chi_j (T) = \int {\cal D} z {\cal D} \bar{z} {\cal D} \psi {\cal D} \bar{\psi} \ \exp\left[ \ii j \int_0^T \dd t \left( \ii {\dot{z} \bar{z} - z\dot{\bar{z}} \over 1 + z\bar{z}} + {1 -z\bar{z} \over 1 + z \bar{z}} + {2\ii \psi \bar{\psi} \over (1 +z\bar{z})^2} \right) \right] \end{equation} with periodic boundary conditions. The Lie-derivative condition (\ref{InvariantMetric}) is satisfied for $H =J_3$. This path integral is given exactly by the WKB-approximation \cite{us,funahashi}. The relevant quantities in the Kirillov formula (\ref{Kirillov}) are \begin{eqnarray}\label{PSInvariance} H- \omega & = & j { 1 -z \bar{z} - \psi \bar{\psi} \over 1 + z \bar{z} + \psi \bar{\psi}} \;, \cr {\cal R}^+ &=& R^+ + \Omega^+ = {1 -z \bar{z} - \psi \bar{\psi} \over 1 +z \bar{z} + \psi \bar{\psi}} \;. \end{eqnarray} Now one can use the Parisi-Sourlas integration formula $$\int \dd z \ \dd \bar{z} \ \dd \psi \ \dd \bar{\psi} F(z \bar{z} + \psi \bar{\psi}) = \pi [F(\infty) - F(0)] $$ which gives \begin{equation}\label{char} \chi_j (T) = {\sin (j+1/2)T \over \sin (T/2)} = \sum_{m = -j}^j \exp[ \ii m T] \:. \end{equation} This is exactly the correct result without an explicit Weyl shift. Also the Weyl formula (\ref{Weyl}) gives the correct result when we use local coordinate charts in the vicinity of the critical points. To get the correct north pole contribution we invert the coordinates $z \rightarrow 1/z, \bar{z} \rightarrow 1/\bar{z}$. This also yields the correct character (\ref{char}): $$ \chi_j (T) = { \exp[-\ii j T ] \over 2 \sin(T/2) } \exp [- \ii T/2 ] + { \exp [-\ii j T (-1) ] \over 2 \sin(-T/2) } \exp[\ii T /2 ] = {\sin(j+1/2)T \over \sin(T/2) } = \sum_{m= -j}^j \exp[\ii m T ] \;. $$ On the other hand, using $\hat{A}$-genus we obtain the result $$ \chi_j (T) = {\sin (j T) \over \sin(T/2) } $$ which is the correct result up to the Weyl shift $j \rightarrow j+1/2$. So we see that in the character formulae we have to use the Todd-genus instead of $\hat{A}$-genus to directly get the correct result. \\ \section{Conclusions} We have considered phase space path integrals with the property that the Hamiltonian generates an isometry of the phase space. Using equivariant cohomology in the loop space we were able to reduce the path integrals to finite dimensional integrals and sums. We also noticed that the results were not uniquely defined because of spectral asymmetry. The choice of regularization yielded equivariant $\hat{A}$- and Todd-classes. We applied localization to the harmonic oscillator and to the quantization of coadjoint orbits. We showed that localization produces correct results for these systems. In addition, we derived Kirillov and Weyl character formulae that produce correct characters for Lie groups without the Weyl shift. We demonstrated this explicitly by evaluating the character for SU(2). The explanation for the Weyl shift was the same as in the case of the Coxeter shift \cite{bbrt} in Chern-Simons theory, the $\eta$-invariant. It would be interesting to apply our formalism to more complicated systems such as loop groups and field theories. Also, it seems possible to use localization and equivariant cohomology to study quantum integrability, generic supersymmetric theories and problems in classical mechanics, as well. \\ {\bf Acknowledgements} \\ We thank Prof. Antti Niemi for initiating this project and for valuable discussions. We also thank A. Alekseev, A. Hietam\"aki and O. Tirkkonen for discussions and comments.
\section{Introduction} The structure functions $G_1$ and $G_2$ describing the spin-dependent part of deep-inelastic scattering were discussed by Feynman in his lectures \cite{Fey} and the definition he used is very appealing due to the simple partonic interpretation of the dimensionless function $g_1(x)$, the only one which survives in the scaling limit for the longitudinal polarization case. However, when one studies the relatively low-$Q^2$ region, the alternative definition of structure functions, proposed by Schwinger a few years later \cite{Sch} appears to be useful as well. In fact, it was recently applied to explain the strong $Q^2$-dependence of the generalized Gerasimov-Drell-Hearn (GDH) sum rule\cite{SoTe} and to clarify a number of the related problems \cite{SoTe2}. In the present paper this approach is applied to the first moment of the $g_1$ structure function entering the fundamental Bjorken sum rule. This quantity was extensively studied in perturbative QCD and the three-loop corrections were calculated exactly \ci{Lar}. As a result, the additional perturbative {\it one-loop} correction, originating from the non-smooth zero-quark-mass limit, is found. The basic point of the approach \cite{SoTe,SoTe2} is the consideration of the structure functions $g_T(x)=g_1(x)+g_2(x)$ and $g_2(x)$ as independent, while $g_1(x)$ (the most familiar one) is expressed as: \begin{equation} \label{def} g_1(x)=g_T(x)-g_2(x) \end{equation} In the resonance region the strong $Q^2-$dependence of $g_2$ due to the BC sum rule naturally provides the similar dependence of the generalized GDH sum rule \ci{SoTe}. At the same time, the BC sum rule in the scaling region is \begin{equation} \label{BC} \int^1_0 g_2(x) dx=0, \end{equation} and, at first sight, absolutely does not contribute to the first moment of $g_1$: \begin{equation} \label{zero} \int^1_0 g_1(x) dx=\int^1_0 g_T(x) dx. \end{equation} However, (\ref{BC}) happens to be valid, if and only if the specific perturbative correction is taken into account. Consequently, this correction should contribute to the r.h.s of (\ref{zero}). The problem arose a couple of years ago due to the paper R. Mertig and W.L. van Neerven{\ci{Mer-vN}, who observed that the BC sum rule is violated in massive perturbative on-shell QCD at one-loop level. However, their result was in contradiction with the QED calculation performed almost 20 years earlier by Wu-Yang Tsai, L.L. DeRaad, Jr. and K.A. Milton \ci{Mil}\footnote{ Probably, this fact was not well known, because the authors of \ci{Mil} used the Schwinger definitions for the spin-dependent structure functions.} (the result in QCD is the same apart from a trivial color factor). The origin of this discreapance was identified in the papers \ci{SoTe2,Mus}. The QED calculation \ci{Mil} treated the fermion mass exactly. The only difference of the QCD case \ci{Mer-vN} is that the quark mass $m$ is just a regulator of collinear singularities, and only the terms contributing to the limit $m\rightarrow 0$ {\it before the integration over $x$} were taken into account. However, this leads to the missing of the finite term at the elastic limit, when the emitted gluon is "soft", \begin{equation} \delta_m^{soft} g_2 (x)=C_F{\alpha_s \over {8 \pi}} \delta_+ (1-x), \end{equation} restoring the validity of the BC sum rule. The same result was obtained by G. Altarelli, B. Lampe, P. Nason and G. Ridolfi \ci{Alt}, who derived the exact mass-dependent formula in QCD, coinciding, up to the mentioned colour factor, with the QED result \cite{Mil}. These authors also established in details the origin of the extra term. This mass correction, due to (\ref{def}) appears in the expression for $g_1$ as well. The important point here is that the similar finite terms for the function $g_T$ are {\it absent} \ci{Mus,Ver}. As a result the full correction to the $g_1$ is just \begin{equation} \delta_m^{soft} g_1 (x) =-\delta_m^{soft} g_2 (x)=C_F{\alpha_s \over {8 \pi}} \delta_+ (1-x). \end{equation} This contribution is especially important in the case of the first moment. The partial conservation of the non-singlet axial current leads to the zero anomalous dimension and to the manifestation of this contribution at the leading approximation. As a result, the correction to the Bjorken sum rule is changed: \begin{equation} \delta_m^{soft} \int_0^1 g_1 (x) dx=-C_F{\alpha_s \over {8 \pi}}, \end{equation} and the full expression for the coefficient function can be written as: \begin{equation} \int_0^1 g_1 (x) dx = {1 \over 2}(1-{3 \over 4}C_F{\alpha_s \over {\pi}}|_{m \equiv 0}-{1 \over 4}C_F{\alpha_s \over {\pi}}|_{m \to 0}^{soft}). \end{equation} This is a principal result of this paper. However, its importance requires further investigation. While such a correction is due to the integration in the region of the "soft" emitted gluons ($x \sim 1$), there is another source of the finite mass-to-zero contribution due to the collinear gluons. In the complete analogy to the $\delta_+ (1-x)$ one get $\delta_+ (\theta^2)$, where $\theta$ is the gluon emission angle in the c.m. frame. As a result, there is a finite helicity-flip cross-section in the zero-mass limit. This effect in QED was discovered by Lee and Nauenberg in their celebrated paper \ci{Lee} and was extensively studied and applied recently \ci{Ein2,Seh}. The main quantitative result is the fermion helicity-flip probability (analogous to the GLAP kernel, except this is finite rather than logarithmic contribution), whose straightforward generalization to the QCD case is: \begin{equation} \label{P+-} P_{+-}(x)= C_F{\alpha_s \over {2 \pi}} (1-x). \end{equation} There is a similar effect in the helicity-non-flip cross section as well. This is easily recovered by, say, the expansion of the numerator of (8) in \ci{Seh}: \begin{equation} {\theta^2 d \theta^2 \over {(\theta^2+m^2/E^2)^2}} ={1 \over {\theta^2+m^2/E^2}}-{m^2/E^2 \over {(\theta^2+m^2/E^2)^2}} \end{equation} Keeping $O(m^2)$ one immediately get the finite correction to the standard logarithmic term, whose $x$-dependence is exactly the same: \begin{equation} \label{P++} P_{++}(x)= - C_F{\alpha_s \over {2 \pi}} \cdot {1+x^2\over {1-x}}. \end{equation} The spin-averaged correction is just \begin{equation} \label{P} P_a(x)= P_{++}(x)+P_{+-}(x)= - C_F{\alpha_s \over { \pi}} \cdot {x\over {1-x}} \end{equation} and exactly coincide with the earlier result of Baier, Fadin and Khoze \ci{Baier}. The agreement with this paper before the angular integration was proved in \cite{Seh}, although the finite correction to the helicity-conserving kernel was not presented explicitly. Let us consider the longitudinally polarized quark. Its density matrix can be represented as a difference of the density matrices with positive and negative helicities. The $g_1$ structure function is proportional to: \begin{equation} g_1(x) \sim \sigma_{++}-\sigma_{+-}-\sigma_{-+}+\sigma_{--}. \end{equation} Taking into account that $\sigma_{--}=\sigma_{++}, \ \sigma_{+-}=\sigma_{-+}$, and that "collinear" contribution results in the substitutions $\sigma_{++} \to \sigma_{++}(1+P_{++})+\sigma_{+-} P_{+-}, \ \sigma_{+-} \to \sigma_{+-}(1+P_{++})+\sigma_{++} P_{+-}$, one should get for the "collinear" correction to $g_1$ \begin{equation} \label{g1c} g_1(x)^{coll}={1\over 2}(\delta (1-x)+P_{++}(x)-P_{+-}(x))={1\over 2} [\delta (1-x)-C_F{\alpha_s \over { \pi}}({1\over {1-x}}+x)], \end{equation} where the Born term is kept to make the normalization clear. Passing to the calculation of the first moment one meet the infrared singularity at $x=1$. It is very important, that the virtual contribution, providing its cancelation, is proportional to the same combination $(log(E^2/m^2)-1)$, appearing in the expression for the GLAP kernel and its correction $P_{++}$. As a result, the $"+"$ prescription should be applicable to $P_{++}$ in complete similarity to the logarithmic term. \begin{equation} \label{P+++} P_{++}(x)= - C_F{\alpha_s \over {2 \pi}} ({1+x^2\over {1-x}})_+. \end{equation} The first moment of $P_{++}$ is then zero and the correction to the Bjorken sum rule is completely determined by the helicity-flip kernel: \begin{equation} \int_0^1 g_1(x)^{coll} dx = {1\over 2}(1-\int_0^1 dx P_{+-}(x))={1\over 2} (1-C_F{\alpha_s \over {4 \pi}}). \end{equation} Finally, the one-loop correction to Bjorken sum rule for the quark is: \begin{equation} \label{tot} \int_0^1 g_1(x) dx = {1\over 2} [1-C_F{\alpha_s \over {4 \pi}}(3(m \equiv 0)+1(soft)+1(coll))], \end{equation} One may worry, is it really possible just to add "zero-mass", "soft" and "collinear" terms. The explicit calculation of $g_{1+2}$ and $g_2$ on mass shell keeping mass exactly \cite{Ver} confirm this naive derivation. The "collinear" contribution to $g_1$ reproduce (\ref{g1c}) while the total correction is just $-5C_F/4 \pi$, like (\ref{tot}). The same value was obtained in \ci{Mer-vN}, (where, however, soft correction was not present) and was interpreted as a manifestation of the regularization dependence. Let us turn to the physical applications of these results. The status of the soft and collinear contributions is then quite a different. The collinear contributions correspond to the integration in the region of the low transverse momenta ($k_T \leq m_q^2$). For the light quarks because of the confinement the quark mass should be replaced by the pion one \cite{GorIo}. This contribution is normally excluded by the cuts when the experimental data are obtained. The situation here is quite analogous to the anomalous gluon contribution to the $g_1$ structure function \cite{EST}. Let us note also in this connection that this naturally explains the results obtained in \cite{Seh2}; namely, that the spin-flip collinear effects contribute to the "normal" piece of the spin-dependent photon spin structure function. This is due to the fact, that these effects are related to the low transverse momenta of scattered quarks. For the heavy quarks, however, this correction should be taken into account. From the other side, there are no reasons to exclude "soft" piece for both heavy and light quarks. It corresponds to the integration over x when $1-x \sim m_q^2/Q^2$. Again, the confinement effects will change quark mass to the pion one. In principle, one can not exclude that the accurate treatment of quark condensates instead of quark mass term may change the actual value of the correction. Anyway, this $x$ due to the standard convolution formula \begin {equation} g_{hadron}(x_B) = \int_{x_B}^1 dx \Delta q(x) g_{quark}(x/x_B) \end{equation} (where $ \Delta q(x)$ is a spin-dependent quark distribution) do not correspond to the particular value of the observed $x_B$. It is interesting, that for the $g_2$ structure function the soft and collinear contributions cancel each other \cite{Ver}. As a result, the Burkhardt-Cottingham sum rule is valid either if one put quark mass equal to zero from the very beginning, or if one take into account all the effects, surviving the zero-mass limit (like for the heavy quarks). If only the soft contributions is taken into account (like for the light quarks), the BC sum rule is violated to the same extent, up to a sign, as was reported by Mertig and van Neerven. This sign difference is due to the fact, that in \ci{Mer-vN} the "collinear' contribution was taken into account (it is just the compensating terms of order $q^2/m^2$ mentioned at p.491 of this reference), contrary to the "soft" one. Finally, inserting the experimentally known expressions for the first moment of the spin-dependent distributions one get the total one-loop correction to $g_1^{p-n}$ \begin {equation} \label{fin} \int_0^1 g_1^{p-n}(x) dx={g_A\over 6} (1-C_F{\alpha_s \over {\pi}}) \end{equation} Let us discuss briefly the experimental situation (see \cite{ALE} and ref. therein). The multiplication of the 1-loop correction to the Bjorken sum rule by the light-quark factor $4/3$ significantly improve the correspondence of the result with the most exact data of E143 experiment at SLAC. Namely, one get the value $1.64 \pm 0.09$ instead of the standard one $1.72 \pm 0.09$. If one take into account the power corrections calculated in the framework of QCD sum rules method results in the value $1.54 \pm 0.017$ instead of $1.62 \pm 0.017$. One should compare this with the most recent experimental value $1.51 \pm 0.013$. Although all the theoretical values coincide with the experimental one within the errors, the mass correction make this agreement for the central values much better. I especially would like to stress the perfect agreement of the QCD sum rules calculation (note that it was obtained with only massless perturbative corrections taken into account, when the agreement is substantially worse). As it is well known, the correction to the Bjorken sum rule coincide with that for the Gross-Llewellynn Smith (GLS) sum rule \ci{Lar,Kat}. Although in the massive case one can not simply transform the relevant diagrams to each other (the $\gamma^5$ moving change the signs of the mass term \footnote{This effect spoils, in principle, the arguments of \ci{Mer-vN}, who used the correction to GLS sum rule calculated earlier to get the answer for Bjorken sum rule.}), $O(m^2)$ terms in the numerator coming from the trace are not responsible for the extra "soft" term in $g_2$ (and, consequently, $g_1$), as it was first mentioned in \cite{Alt}. As a result, the coefficient function for the light quarks receives the same contribution: \begin {equation} \label{GLS} \int_0^1 F_3(x) dx = 3(1-C_F{\alpha_s \over \pi}). \end{equation} This relation, however, requires an additional check by the straightforward calculation, because one can not exclude, in principle, the appearance of such corrections coming from another terms. This work is now in progress \cite{Ver2}. However, the extra factor $4/3$ already completely remove the discreapance \cite{Sidor} of the data with perturbative QCD for all $Q^2$. Note that such a 'compensating' factor fully respects the qualitative nature of this discreapance: the experimental data as a function of $Q^2$ are going above the theoretical curve and the difference is decreasing for higher $Q^2$. The corrections to the Bjorken and GLS sum rules are in turn related to the correction to the $e^+ e^-$-annihilation total cross section. This is due to the famous Crewther relation \cite{Crewter}, studied recently for high orders of perturbation theory\cite{Kat}. However, the zero mass limit in this case is known to be smooth\cite{PSF}. It seems that there is no contradiction here. The Crewther relation emerges due to the non-renormalization of axial anomaly, which is not directly related to the mass contributions studied here. In conclusion, the new perturbative QCD correction to the partonic sum rules is found. It comes from the quark mass contribution, having the non-zero limit, when the quark mass tends to zero. This correction seems to improve the agreement with the experimental data for Bjorken sum rule. Taking into account the similar correction removes the deficit of Gross-LLewellyn Smith sum rule. Although the manifestation of these correction requires further investigation, one may conclude that the study of mass effects (even in a zero mass limit) in hard processes (in particular, in the partonic sum rules) seems to be very interesting. I am thankful to A.V. Efremov, J. Ho\'rej\'si, E.A. Kuraev and S.V. Mikhailov for stimulating discussions and valuable comments. \begin{thebibliography}} \newcommand{\eb{99} \bi{Fey} R.P.Feynman {\it Photon-Hadron Interactions}, (Benjamin, Reading, MA, 1972). \bi{Sch} J.Schwinger, Proc. Natl. Acad. Sci. U.S.A. {\bf 72}, (1975) 1559. \bi{SoTe} J.Soffer and O.Teryaev, Phys. Rev. Lett. {\bf 70}, (1993) 3373. \bi{SoTe2} J.Soffer and O.Teryaev, Phys. Rev. {\bf D51}, (1995) 25. \bi{Lar} S.A. Larin, J.A.M. Vermaseren, Phys. Lett {\bf B259}, (1991) 345. \bi{Mer-vN} R. Mertig and W.L. van Neerven, Z. Phys.{\bf C60}, (1993) 489. \bi{Mil} Wu-Yang Tsai, L.L. DeRaad, Jr. and K.A. Milton, Phys. Rev. {\bf D11}, (1975) 3537. \bi{Mus} I.V. Musatov, O.V. Teryaev, CEBAF Report TH-94-25(unpublished). \bi{Alt} G. Altarelli, B. Lampe, P. Nason and G. Ridolfi, Phys. Lett {\bf B334}, (1994) 187. \bi{Ver} I.V. Musatov, O.V. Teryaev and O.L. Veretin, in preparation. \bi{Lee} T.D. Lee and M. Nauenberg, Phys. Rev. {\bf B133}, (1964) 1549. \bi{Ein2} H.F. Contopanagos and M.B. Einhorn, Nucl. Phys {\bf B377}, (1992) 20 \bi{Seh} B. Falk and L.M. Sehgal, Phys. Lett {\bf B325}, (1994) 509. \bi{Baier} V.N. Baier, V.S. Fadin and V.A. Khoze, Nucl. Phys. {\bf B65}, (1973) 381. \bi{GorIo} A.S. Gorskii, B.L. Ioffe, A.Yu. Khodzhamirian, Phys. Lett. {\bf B227}, (1989) 474. \bi{EST} A.V.Efremov, J.Soffer, O.V.Teryaev, Nucl.Phys. {\bf B346}, (1990) 97. \bi{Seh2} A. Freund and L.M. Sehgal, Phys. Lett {\bf B341}, (1994) 90. \bi{ALE} M. Anselmino, A.V. Efremov and E. Leader, CERN Preprint CERN-TH.7216/94 (to be published in Physics Reports). \bi{Kat} D.J. Broadhurst and A.L. Kataev, Phys. Lett {\bf B315}, (1993) 179 \bi{Sidor} A.L. Kataev and A.V. Sidorov, Phys. Lett. {\bf B331}, (1994) 179 \bi{Ver2} O.V. Teryaev, O.L. Veretin, in progress. \bi{Crewter} R.J. Crewther, Phys. Rev. Lett. {\bf 28}, (1972) 345. \bi{PSF} J.Schwinger, {\it Particles, Sources and Fields} (Addison-Wesley, 1989) v. 3, p.99. \eb \end{document}
\section{Introduction} In the period leading up to the start of the HERA program a substantial amount of work was done on radiative corrections to observables in deep-inelastic scattering (DIS) of electrons off protons (see e.g. \cite{HERAproc} for an overview). These corrections are substantial enough for both HERA experiments to correct for them. Part of the upcoming LEP2 physics programme involves the measurement of the photon structure function $F_2^{\gamma}(x,Q^2)$. This structure function is extracted from deep-inelastic electron--photon scattering in the reaction ${\mathrm{e}}^+{\mathrm{e}}^- \rightarrow {\mathrm{e}}^+{\mathrm{e}}^- X$, where one of the leptons escapes undetected down the beam pipe, while the other is measured at rather large angle. It is therefore important to study radiative corrections to deep-inelastic electron--photon scattering. This is our purpose in this paper. Radiative corrections to ${\mathrm{e}}^+{\mathrm{e}}^- \rightarrow {\mathrm{e}}^+{\mathrm{e}}^- X$ have been calculated for $X$ a (pseudo) scalar particle \cite{Defrise,Neerven,Landro} and $X = \mu^+\mu^-$ \cite{Landro,Berends}. It was found that they are very small (on the percent level) in the no-tag case, when neither the electron nor the positron is being measured. In such a kinematic configuration, when the momentum transfer between the incident and outgoing electron (or positron) is small, the vertex- and bremsstrahlung contributions effectively cancel each other, leaving a small correction dominated by vacuum polarization \cite{Neerven,Landro}. This implies that for the equivalent photon spectrum, which is essential to relate ${\mathrm{e}}^{\pm}\gamma$ with ${\mathrm{e}}^+{\mathrm{e}}^-$ reactions, corrections are small. In contrast, radiative corrections can be sizeable in the case where one of the leptons scatters at a large angle - single tag - and one studies differential cross sections which depend strongly on the energy and angle of the tagged electron (or positron). Surprisingly, no calculations of the size of radiative corrections for inclusive deep-inelastic electron--photon scattering, i.e. $e\gamma \rightarrow e X$ with both large $Q^2$ (the absolute value of the transferred momentum squared) and $W$ (the mass of state $X$), have been performed yet. Correspondingly, in experimental analyses of the (hadronic) photon structure function $F_2^\gamma(x,Q^2)$ radiative corrections have so far not been included (as usual, $x=Q^2/(Q^2+W^2)$). They are usually assumed to be negligible. Recently, the AMY collaboration \cite{AMY} estimated the size of radiative corrections by comparing a Monte-Carlo event generator based on the full cross section formula for ${\mathrm{e}}^+{\mathrm{e}}^- \rightarrow {\mathrm{e}}^+ {\mathrm{e}}^- \gamma\mu^+\mu^-$ \cite{Berends} (with the muon mass and electric charge changed to correspond with quark-antiquark pair production) with a generator for ${\mathrm{e}}^+{\mathrm{e}}^- \rightarrow {\mathrm{e}}^+{\mathrm{e}}^- {\mathrm{q}\bar{\mathrm{q}}}$ where the cross section for ${\mathrm{e}} \gamma \rightarrow {\mathrm{e}} {\mathrm{q}\bar{\mathrm{q}}}$ with a real photon was convoluted with the equivalent photon spectrum. A (positive) correction of order $10$\% for the visible $x$ distribution was found, which, however, cancelled effectively against the correction due to a non-vanishing target-photon mass. Hence no net correction was applied. Nevertheless, it is important to understand both corrections separately, and their behavior as a function of the kinematic variables chosen and phase space. Moreover, the hadronic structure of the photon must not be neglected. Here we therefore estimate the size of the radiative corrections to inclusive deep-inelastic ${\mathrm{e}}\gamma$ scattering, using the full photon structure function. In the next section we describe the relevant kinematics and formalism and in section 3 we present results. \section{Formalism} We consider the $O(\alpha)$ corrections to deep-inelastic scattering (DIS) of electrons on (quasi-real) photons: \begin{equation} {\mathrm{e}}(l) + \gamma(p) \rightarrow {\mathrm{e}}(l') + \gamma(k) + X(p_X) \ . \label{DISreact} \end{equation} This process is depicted in Fig.1, in which we indicate all momentum labels. The target photon $\gamma(p)$ is part of the flux of equivalent photons around the non-tagged lepton. We assume that this flux has a momentum density given by the Weizs\"{a}cker-Williams expression \begin{equation} f_{\gamma/{\mathrm{e}}}(z) = \frac{\alpha}{2\pi} \left\{ \frac{1+(1-z)^2}{z}\, \ln \frac{P^2_{max}}{P^2_{min}} - 2 m_e^2 z(\frac{1}{P^2_{min}} -\frac{1}{P^2_{max}})\right\} \label{fww} \end{equation} where $P^2_{min} = (z^2 m_e^2)/(1-z)$ and $P^2_{max} = (1-z) \left(E_b \theta_{max}\right)^2$. Here $z$ is the longitudinal momentum fraction of the target photon with respect to its parent lepton, $E_b=\sqrt{s}/2$ is the lepton beam energy, $\theta_{max}$ is the anti-tag \footnote{i.e. all events in which the parent lepton scatters at an angle larger than $\theta_{max}$ are rejected.} angle and $P^2 = -p^2$. In the following we put $P^2 = 0$ and neglect electron masses everywhere except in (\ref{fww}). Moreover we substitute $P^2_{max}$ by $P^2_{max}+P^2_{min}$ so that we can easily extend the $z$ range to 1, see \cite{FMNR}. \vglue 6.1cm \vbox{\special{psfile=fig1.ps angle=0 hscale=60 vscale=60 hoffset=40 voffset=-180}} {\tenrm\baselineskip=12pt \noindent Figure 1: Photon bremsstrahlung from the tagged lepton line in deep-inelastic scattering off an equivalent photon.} \vglue 0.8cm The DIS variables can be defined either from the leptonic or the hadronic momenta: \begin{equation} \begin{array}{rclrcl} q_l & = & l-l' & q_h & = & p_X - p = q_l - k\\ W_l^2 & = & (p+q_l)^2 \qquad & W_h^2 & = & (p+q_h)^2 = p_X^2\\ Q_l^2 & = & - q_l^2 & Q_h^2 & = & - q_h^2\\ x_l & = & Q_l^2/2p \cdot q_l & x_h & = & Q_h^2/2p \cdot q_h\\ y_l & = & p \cdot q_l/p \cdot l & y_h & = & p \cdot q_h/p \cdot l \end{array} \label{DISvar} \end{equation} Note that both $Q_l^2 = x_l y_l s_{\re \gamma}$ and $Q_h^2 = x_h y_h s_{\re \gamma}$, where $s_{\re \gamma} = (p+l)^2$, but that leptonic and hadronic variables agree only for nonradiative events, i.e. if $k=0$. We will see that the size of the corrections strongly depends on which set of variables are used in the measurement. In practice one determines $Q^2$ from the tagged lepton, and $x$ from the (visible) hadronic energy $W_h$. The Born cross section (i.e. no $\gamma(k)$ in (\ref{DISreact})) is given by \begin{equation} \frac{{\mathrm{d}}^2\sigma^{\mathrm{B}}}{{\mathrm{d}} x {\mathrm{d}} Q^2} = f^{\mathrm{B}}(x,Q^2,s) \ , \label{sigBorn} \end{equation} where \begin{eqnarray} f^{\mathrm{B}}(x,Q^2,s) &= & \frac{2\pi\alpha^2}{x Q^4}\, F_2(x,Q^2) \\ & \times & \int_{z_{min}(x,Q^2,s)}^1 {\mathrm{d}} z\, f_{\gamma/{\mathrm{e}}}(z)\, Y_+\left(\frac{Q^2}{x z s} \right) \, \left\{ 1 + R\left(x,Q^2,\frac{Q^2}{x z s} \right) \right\} \label{Borncross} \ . \end{eqnarray} Here we have defined $z_{min}(x,Q^2,s) = Q^2/xs$, $ Y_+(y) = 1 + (1-y)^2$ and \begin{equation} R(x,Q^2,y) = \frac{- y^2}{1 +(1-y)^2} \ \frac{F_L(x,Q^2)}{F_2(x,Q^2)} \ . \label{Rdef} \end{equation} $F_{2,L}$ are the photon structure functions (we have dropped the superscript $\gamma$). The above form eq.~(\ref{Borncross}) is useful because $F_2$ can be factored out of the $z$ integration. For comparison of ${\mathrm{e}} \gamma$ with ${\mathrm{e}} {\mathrm{p}}$ scattering we will also give the cross section in terms of $x$ and $y$ \begin{equation} \frac{{\mathrm{d}}^2\sigma^{\mathrm{B}}}{{\mathrm{d}} x {\mathrm{d}} y} = g^{\mathrm{B}}(x,y,s) \ , \label{nsigBorn} \end{equation} where \begin{equation} g^{\mathrm{B}}(x,y,s) = \frac{2 \pi\alpha^2 \,Y_+(y)}{x^2 y^2 s} \int_{\epsilon(x,y,s)}^1 \frac{{\mathrm{d}} z}{z} \, f_{\gamma/{\mathrm{e}}}(z)\, F_2(x,xyzs) \left\{1 + R(x,xyzs,y) \right\} \ . \label{newBorn} \end{equation} The lower limit $\epsilon = W^2_{\mathrm{min}}/(1-x)ys$ on the $z$-integration in eq.~(\ref{newBorn}) arises if a lower cut is applied to the invariant hadronic mass $W$. At the Born level expressions (\ref{sigBorn}) and (\ref{nsigBorn}) are equivalent and valid for both sets of variables in (\ref{DISvar}). For the ${\mathrm{e}}$p case the full electroweak corrections have been calculated for both neutral current reactions \cite{Dubna,Wurzburg}, including elastic nucleon scattering \cite{Akhundov}, and charged-current reactions \cite{Bohm}. It is well-known that the leading logarithmic approximation (LLA) (in $\ln(Q^2/m_e^2)$) reproduces the exact results to within a few percent however \cite{Consoli, Spiesberger}. Here we will therefore work within this approximation and neglect furthermore $Z$ exchange. We consider first the radiative corrections to the differential cross section in (\ref{nsigBorn}) in terms of $x_l$ and $y_l$. The $O(\alpha)$ corrections in LLA arise from collinear and soft bremsstrahlung from the initial and final electron that couple to the ``probing photon" and from the Compton process\footnote{As stated above, radiative effects to the Weizs\"{a}cker-Williams spectrum are small \cite{Landro} and we therefore neglect them.}. The latter corresponds to the case in Fig.1 where the exchanged photon $\gamma(q)$ is quite soft, but $\gamma(k)$ is radiated at a wide angle causing the lepton $e(l')$ to be tagged. The bremsstrahlung terms are given by \begin{equation} \frac{{\mathrm{d}}^2\sigma^{\mathrm{Br}}}{{\mathrm{d}} x {\mathrm{d}} y} = \int_0^1 {\mathrm{d}} x_i \, D_{e/e}(x_i,Q^2) \left\{ \Theta\left( x_i - x_i^0\right) J(x_1,x_2) g^{\mathrm{B}}(\hat{x},\hat{y},\hat{s}) - g^{\mathrm{B}}(x,y,s) \right\} \label{bremscross} \end{equation} where $\hat{s} = x_1 s$, $\hat{x} = x x_1 y/(x_1 x_2 + y -1)$, $\hat{y} = y x/x_2 \hat{x}$, $J(x_1,x_2) = y/x_1 x_2^2 \hat{y}$, $x_1^0 = (1-y)/(1-x y)$ and $x_2^0 = x y +1-y$, and \begin{equation} D_{e/e}(x,Q^2) = \frac{\alpha}{2\pi}\ln\left(\frac{Q^2}{m_e^2}\right) \, \frac{1+x^2}{1-x}\ . \end{equation} In eq. (\ref{bremscross}) we have supressed the subscript $l$ on $x$ and $y$ for clarity. Initial-state radiation corresponds to $x_i = x_1$ and $x_2=1$ in eq.~(\ref{bremscross}) and vice versa for final-state radiation. The Compton contribution is given by \begin{equation} \frac{{\mathrm{d}}^2\sigma^{\mathrm{C}}}{{\mathrm{d}} x_l {\mathrm{d}} y_l} = \int_{\epsilon}^1 {\mathrm{d}} z \, f_{\gamma/{\mathrm{e}}}(z)\, h^{{\mathrm{C}}}(x_l,y_l,x_ly_l zs,zs) \ , \label{Compton} \end{equation} where \begin{eqnarray} h^{{\mathrm{C}}}(x,y,Q^2,s) &=& \frac{\alpha^3}{x^2(1-y)s}\ Y_+(y)\ \ln\frac{Q^2}{M^2} \int_x^1 \frac{{\mathrm{d}} v}{v} \left[ 1 + \left(1 - \frac{x}{v}\right)^2 \right]\nonumber\\ &\times& F_2(v,Q^2) (1+R(v,Q^2,y)) \ . \label{hCdef} \end{eqnarray} The logarithm $\ln(Q^2/M^2)$ is a result of the absorption of the collinear singularity from the quark to photon splitting by renormalizing the photon density in the quark at scale $M$. We take $M$ here to be the proton mass \cite{Dubna}. Next we discuss the radiative corrections to the differential cross section expressed in hadronic variables. Now there are, in accordance with the KLN theorem \cite{KLN}, in LLA approximation neither corrections from final state radiation, nor from the Compton process because these process do not affect the kinematic variables. We find that the correction due to initial state radiation can simply be expressed as \begin{eqnarray} \frac{{\mathrm{d}}^2\sigma^{\rm corr}}{{\mathrm{d}} x_h {\mathrm{d}} Q_h^2}& = & \frac{2 \pi\alpha^2}{x_h Q_h^4}{\Bigg\{} F_2(x_h,Q_h^2) \, \int_{z_{min}(x_h,Q_h^2,s)}^1 {\mathrm{d}} z\, f_{\gamma/{\mathrm{e}}}(z)\, g_h\left( x_h,Q_h^2,\frac{Q_h^2}{z x_h s}\right) \nonumber \\ & + & F_L(x_h,Q_h^2) \, \int_{z_{min}(x_h,Q_h^2,s)}^1 {\mathrm{d}} z\, f_{\gamma/{\mathrm{e}}}(z)\, h_h\left( x_h,Q_h^2,\frac{Q_h^2}{z x_h s}\right) {\Bigg\}} \label{hadcorr} \end{eqnarray} where \begin{equation} g_h(x,Q^2,y) = \frac{\alpha}{\pi}\ \ln \frac{Q^2}{m_e^2}\, \left\{ Y_+(y) \ln (1-y) + y\left(1 - \frac{y}{2}\right) \ln y + y\left(1 - \frac{y}{4} \right) \right\} \end{equation} and \begin{equation} h_h(x,Q^2,y) = \frac{\alpha}{\pi}\ \ln \frac{Q^2}{m_e^2}\, \left\{ -y^2 \ln (1-y) + \frac{y^2}{2}\ln y - \frac{y}{2}\left(1 - \frac{y}{2} \right) \right\} \end{equation} The corrections are large at large $y$ (soft region), and can in fact be resummed by simple exponentiation, with the result \begin{eqnarray} g_h(x,Q^2,y) & = & Y_+(y)\ \left\{ \exp \left[ \frac{\alpha}{\pi}\ \left( \ln \frac{Q^2}{m_e^2} - 1 \right) \ln(1-y) \right] - 1 \right\} \nonumber\\ & & \quad + \frac{\alpha}{\pi}\ \ln \frac{Q^2}{m_e^2}\, \left\{ y\left(1 - \frac{y}{2}\right) \ln y + y\left(1 - \frac{y}{4} \right) \right\} \label{resummed2} \end{eqnarray} and \begin{eqnarray} h_h(x,Q^2,y) & = & -y^2\ \left\{ \exp \left[ \frac{\alpha}{\pi}\ \left( \ln \frac{Q^2}{m_e^2} - 1 \right) \ln(1-y) \right] - 1 \right\} \nonumber\\ & & \quad + \frac{\alpha}{\pi}\ \ln \frac{Q^2}{m_e^2}\, \left\{ \left(\frac{y^2}{2}\right) \ln y -\frac{y}{2}\left(1 - \frac{y}{2} \right) \right\} \label{resummedL} \end{eqnarray} In the next section we use the formulae listed in the above to estimate the size of the corrections. \section{Results} Here we study the radiative correction to deep-inelastic electron--photon scattering numerically. For the results presented below we use $\sqrt{s} = 175\,{\rm GeV}$ and $W_{min} = 2\,{\rm GeV}$. For the parton densities in the photon we use set 1 of \cite{SaS}. This set has an already low minimum $Q^2$ of $Q_0^2 =0.36\, {\rm GeV}^2$. However in radiative events even lower values of $Q^2$ contribute. We therefore extrapolate below this value by \begin{eqnarray} F_2^{\gamma}(x,Q^2<Q_0^2)& = &F_2^{\gamma}(x,Q_0^2)\left(\frac{Q^2}{Q_0^2}\right)^2+ \left(1-\frac{Q^2}{Q_0^2}\right) \frac{Q^2 (1-x)}{112\,{\rm GeV^2}} \nonumber \\ &\times&\left(0.211 (\frac{W^2}{\rm GeV^2})^{0.08} + 0.297 (\frac{W^2}{\rm GeV^2})^{-0.45}\right), \label{extrapol} \end{eqnarray} (this expression correctly approaches the $\gamma\gamma$ total cross section in the small $Q^2$ limit) which vanishes as $Q^2\rightarrow 0$, as required by gauge invariance. Furthermore we have neglected $F_L^{\gamma}$, i.e. we put $R$ in (\ref{Borncross}) and (\ref{newBorn}) to zero. We found that its inclusion has a negligible effect. \vglue 0.5cm \setlength{\unitlength}{0.240900pt} \ifx\plotpoint\undefined\newsavebox{\plotpoint}\fi \sbox{\plotpoint}{\rule[-0.200pt]{0.400pt}{0.400pt}}% \begin{picture}(1200,1259)(0,0) \font\gnuplot=cmr10 at 10pt \gnuplot \sbox{\plotpoint}{\rule[-0.200pt]{0.400pt}{0.400pt}}% \put(176.0,562.0){\rule[-0.200pt]{231.264pt}{0.400pt}} \put(176.0,113.0){\rule[-0.200pt]{0.400pt}{270.531pt}} \put(176.0,113.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,113){\makebox(0,0)[r]{-0.8}} \put(1116.0,113.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,225.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,225){\makebox(0,0)[r]{-0.6}} \put(1116.0,225.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,338.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,338){\makebox(0,0)[r]{-0.4}} \put(1116.0,338.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,450.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,450){\makebox(0,0)[r]{-0.2}} \put(1116.0,450.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,562.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,562){\makebox(0,0)[r]{0}} \put(1116.0,562.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,675.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,675){\makebox(0,0)[r]{0.2}} \put(1116.0,675.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,787.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,787){\makebox(0,0)[r]{0.4}} \put(1116.0,787.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,899.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,899){\makebox(0,0)[r]{0.6}} \put(1116.0,899.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,1011.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,1011){\makebox(0,0)[r]{0.8}} \put(1116.0,1011.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,1124.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,1124){\makebox(0,0)[r]{1}} \put(1116.0,1124.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,1236.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,1236){\makebox(0,0)[r]{1.2}} \put(1116.0,1236.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(176,68){\makebox(0,0){0}} \put(176.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(283.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(283,68){\makebox(0,0){0.1}} \put(283.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(389.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(389,68){\makebox(0,0){0.2}} \put(389.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(496.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(496,68){\makebox(0,0){0.3}} \put(496.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(603.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(603,68){\makebox(0,0){0.4}} \put(603.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(709.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(709,68){\makebox(0,0){0.5}} \put(709.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(816.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(816,68){\makebox(0,0){0.6}} \put(816.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(923.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(923,68){\makebox(0,0){0.7}} \put(923.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1029.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1029,68){\makebox(0,0){0.8}} \put(1029.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1136.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1136,68){\makebox(0,0){0.9}} \put(1136.0,1216.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(176.0,113.0){\rule[-0.200pt]{231.264pt}{0.400pt}} \put(1136.0,113.0){\rule[-0.200pt]{0.400pt}{270.531pt}} \put(176.0,1236.0){\rule[-0.200pt]{231.264pt}{0.400pt}} \put(656,23){\makebox(0,0){$y_l$}} \put(176.0,113.0){\rule[-0.200pt]{0.400pt}{270.531pt}} \put(187,423){\usebox{\plotpoint}} \multiput(187.58,423.00)(0.498,0.991){81}{\rule{0.120pt}{0.890pt}} \multiput(186.17,423.00)(42.000,81.152){2}{\rule{0.400pt}{0.445pt}} \multiput(229.00,506.58)(0.751,0.498){69}{\rule{0.700pt}{0.120pt}} \multiput(229.00,505.17)(52.547,36.000){2}{\rule{0.350pt}{0.400pt}} \multiput(283.00,542.58)(1.299,0.498){79}{\rule{1.134pt}{0.120pt}} \multiput(283.00,541.17)(103.646,41.000){2}{\rule{0.567pt}{0.400pt}} \multiput(389.00,583.58)(1.928,0.497){53}{\rule{1.629pt}{0.120pt}} \multiput(389.00,582.17)(103.620,28.000){2}{\rule{0.814pt}{0.400pt}} \multiput(496.00,611.58)(2.254,0.496){45}{\rule{1.883pt}{0.120pt}} \multiput(496.00,610.17)(103.091,24.000){2}{\rule{0.942pt}{0.400pt}} \multiput(603.00,635.58)(1.781,0.497){57}{\rule{1.513pt}{0.120pt}} \multiput(603.00,634.17)(102.859,30.000){2}{\rule{0.757pt}{0.400pt}} \multiput(709.00,665.58)(1.632,0.497){63}{\rule{1.397pt}{0.120pt}} \multiput(709.00,664.17)(104.101,33.000){2}{\rule{0.698pt}{0.400pt}} \multiput(816.00,698.58)(1.012,0.498){103}{\rule{0.908pt}{0.120pt}} \multiput(816.00,697.17)(105.116,53.000){2}{\rule{0.454pt}{0.400pt}} \multiput(923.00,751.58)(0.552,0.498){93}{\rule{0.542pt}{0.120pt}} \multiput(923.00,750.17)(51.876,48.000){2}{\rule{0.271pt}{0.400pt}} \multiput(976.58,799.00)(0.498,0.547){103}{\rule{0.120pt}{0.538pt}} \multiput(975.17,799.00)(53.000,56.884){2}{\rule{0.400pt}{0.269pt}} \multiput(1029.58,857.00)(0.498,1.096){105}{\rule{0.120pt}{0.974pt}} \multiput(1028.17,857.00)(54.000,115.978){2}{\rule{0.400pt}{0.487pt}} \multiput(1083.58,975.00)(0.498,2.085){103}{\rule{0.120pt}{1.760pt}} \multiput(1082.17,975.00)(53.000,216.346){2}{\rule{0.400pt}{0.880pt}} \put(187,370){\usebox{\plotpoint}} \multiput(187.58,370.00)(0.498,0.931){81}{\rule{0.120pt}{0.843pt}} \multiput(186.17,370.00)(42.000,76.251){2}{\rule{0.400pt}{0.421pt}} \multiput(229.00,448.58)(0.731,0.498){71}{\rule{0.684pt}{0.120pt}} \multiput(229.00,447.17)(52.581,37.000){2}{\rule{0.342pt}{0.400pt}} \multiput(283.00,485.58)(1.331,0.498){77}{\rule{1.160pt}{0.120pt}} \multiput(283.00,484.17)(103.592,40.000){2}{\rule{0.580pt}{0.400pt}} \multiput(389.00,525.58)(2.000,0.497){51}{\rule{1.685pt}{0.120pt}} \multiput(389.00,524.17)(103.502,27.000){2}{\rule{0.843pt}{0.400pt}} \multiput(496.00,552.58)(2.463,0.496){41}{\rule{2.045pt}{0.120pt}} \multiput(496.00,551.17)(102.755,22.000){2}{\rule{1.023pt}{0.400pt}} \multiput(603.00,574.58)(3.174,0.495){31}{\rule{2.594pt}{0.119pt}} \multiput(603.00,573.17)(100.616,17.000){2}{\rule{1.297pt}{0.400pt}} \multiput(709.00,591.58)(2.463,0.496){41}{\rule{2.045pt}{0.120pt}} \multiput(709.00,590.17)(102.755,22.000){2}{\rule{1.023pt}{0.400pt}} \multiput(816.00,613.58)(1.928,0.497){53}{\rule{1.629pt}{0.120pt}} \multiput(816.00,612.17)(103.620,28.000){2}{\rule{0.814pt}{0.400pt}} \multiput(923.00,641.58)(1.490,0.495){33}{\rule{1.278pt}{0.119pt}} \multiput(923.00,640.17)(50.348,18.000){2}{\rule{0.639pt}{0.400pt}} \multiput(976.00,659.58)(0.951,0.497){53}{\rule{0.857pt}{0.120pt}} \multiput(976.00,658.17)(51.221,28.000){2}{\rule{0.429pt}{0.400pt}} \multiput(1029.00,687.58)(0.587,0.498){89}{\rule{0.570pt}{0.120pt}} \multiput(1029.00,686.17)(52.818,46.000){2}{\rule{0.285pt}{0.400pt}} \multiput(1083.58,733.00)(0.498,0.851){103}{\rule{0.120pt}{0.779pt}} \multiput(1082.17,733.00)(53.000,88.383){2}{\rule{0.400pt}{0.390pt}} \put(187,343){\usebox{\plotpoint}} \multiput(187.58,343.00)(0.498,0.919){81}{\rule{0.120pt}{0.833pt}} \multiput(186.17,343.00)(42.000,75.270){2}{\rule{0.400pt}{0.417pt}} \multiput(229.00,420.58)(0.731,0.498){71}{\rule{0.684pt}{0.120pt}} \multiput(229.00,419.17)(52.581,37.000){2}{\rule{0.342pt}{0.400pt}} \multiput(283.00,457.58)(1.299,0.498){79}{\rule{1.134pt}{0.120pt}} \multiput(283.00,456.17)(103.646,41.000){2}{\rule{0.567pt}{0.400pt}} \multiput(389.00,498.58)(2.000,0.497){51}{\rule{1.685pt}{0.120pt}} \multiput(389.00,497.17)(103.502,27.000){2}{\rule{0.843pt}{0.400pt}} \multiput(496.00,525.58)(2.714,0.496){37}{\rule{2.240pt}{0.119pt}} \multiput(496.00,524.17)(102.351,20.000){2}{\rule{1.120pt}{0.400pt}} \multiput(603.00,545.58)(2.994,0.495){33}{\rule{2.456pt}{0.119pt}} \multiput(603.00,544.17)(100.903,18.000){2}{\rule{1.228pt}{0.400pt}} \multiput(709.00,563.58)(3.204,0.495){31}{\rule{2.618pt}{0.119pt}} \multiput(709.00,562.17)(101.567,17.000){2}{\rule{1.309pt}{0.400pt}} \multiput(816.00,580.58)(2.714,0.496){37}{\rule{2.240pt}{0.119pt}} \multiput(816.00,579.17)(102.351,20.000){2}{\rule{1.120pt}{0.400pt}} \multiput(923.00,600.58)(2.083,0.493){23}{\rule{1.731pt}{0.119pt}} \multiput(923.00,599.17)(49.408,13.000){2}{\rule{0.865pt}{0.400pt}} \multiput(976.00,613.58)(1.490,0.495){33}{\rule{1.278pt}{0.119pt}} \multiput(976.00,612.17)(50.348,18.000){2}{\rule{0.639pt}{0.400pt}} \multiput(1029.00,631.58)(0.935,0.497){55}{\rule{0.845pt}{0.120pt}} \multiput(1029.00,630.17)(52.247,29.000){2}{\rule{0.422pt}{0.400pt}} \multiput(1083.58,660.00)(0.498,0.509){103}{\rule{0.120pt}{0.508pt}} \multiput(1082.17,660.00)(53.000,52.947){2}{\rule{0.400pt}{0.254pt}} \put(187,322){\usebox{\plotpoint}} \multiput(187.58,322.00)(0.498,0.919){81}{\rule{0.120pt}{0.833pt}} \multiput(186.17,322.00)(42.000,75.270){2}{\rule{0.400pt}{0.417pt}} \multiput(229.00,399.58)(0.731,0.498){71}{\rule{0.684pt}{0.120pt}} \multiput(229.00,398.17)(52.581,37.000){2}{\rule{0.342pt}{0.400pt}} \multiput(283.00,436.58)(1.299,0.498){79}{\rule{1.134pt}{0.120pt}} \multiput(283.00,435.17)(103.646,41.000){2}{\rule{0.567pt}{0.400pt}} \multiput(389.00,477.58)(2.000,0.497){51}{\rule{1.685pt}{0.120pt}} \multiput(389.00,476.17)(103.502,27.000){2}{\rule{0.843pt}{0.400pt}} \multiput(496.00,504.58)(2.714,0.496){37}{\rule{2.240pt}{0.119pt}} \multiput(496.00,503.17)(102.351,20.000){2}{\rule{1.120pt}{0.400pt}} \multiput(603.00,524.58)(3.378,0.494){29}{\rule{2.750pt}{0.119pt}} \multiput(603.00,523.17)(100.292,16.000){2}{\rule{1.375pt}{0.400pt}} \multiput(709.00,540.58)(3.204,0.495){31}{\rule{2.618pt}{0.119pt}} \multiput(709.00,539.17)(101.567,17.000){2}{\rule{1.309pt}{0.400pt}} \multiput(816.00,557.58)(3.022,0.495){33}{\rule{2.478pt}{0.119pt}} \multiput(816.00,556.17)(101.857,18.000){2}{\rule{1.239pt}{0.400pt}} \multiput(923.00,575.58)(2.477,0.492){19}{\rule{2.027pt}{0.118pt}} \multiput(923.00,574.17)(48.792,11.000){2}{\rule{1.014pt}{0.400pt}} \multiput(976.00,586.58)(1.929,0.494){25}{\rule{1.614pt}{0.119pt}} \multiput(976.00,585.17)(49.649,14.000){2}{\rule{0.807pt}{0.400pt}} \multiput(1029.00,600.58)(1.238,0.496){41}{\rule{1.082pt}{0.120pt}} \multiput(1029.00,599.17)(51.755,22.000){2}{\rule{0.541pt}{0.400pt}} \multiput(1083.00,622.58)(0.680,0.498){75}{\rule{0.644pt}{0.120pt}} \multiput(1083.00,621.17)(51.664,39.000){2}{\rule{0.322pt}{0.400pt}} \put(187,304){\usebox{\plotpoint}} \multiput(187.58,304.00)(0.498,0.919){81}{\rule{0.120pt}{0.833pt}} \multiput(186.17,304.00)(42.000,75.270){2}{\rule{0.400pt}{0.417pt}} \multiput(229.00,381.58)(0.731,0.498){71}{\rule{0.684pt}{0.120pt}} \multiput(229.00,380.17)(52.581,37.000){2}{\rule{0.342pt}{0.400pt}} \multiput(283.00,418.58)(1.299,0.498){79}{\rule{1.134pt}{0.120pt}} \multiput(283.00,417.17)(103.646,41.000){2}{\rule{0.567pt}{0.400pt}} \multiput(389.00,459.58)(2.000,0.497){51}{\rule{1.685pt}{0.120pt}} \multiput(389.00,458.17)(103.502,27.000){2}{\rule{0.843pt}{0.400pt}} \multiput(496.00,486.58)(2.714,0.496){37}{\rule{2.240pt}{0.119pt}} \multiput(496.00,485.17)(102.351,20.000){2}{\rule{1.120pt}{0.400pt}} \multiput(603.00,506.58)(3.174,0.495){31}{\rule{2.594pt}{0.119pt}} \multiput(603.00,505.17)(100.616,17.000){2}{\rule{1.297pt}{0.400pt}} \multiput(709.00,523.58)(3.410,0.494){29}{\rule{2.775pt}{0.119pt}} \multiput(709.00,522.17)(101.240,16.000){2}{\rule{1.388pt}{0.400pt}} \multiput(816.00,539.58)(3.204,0.495){31}{\rule{2.618pt}{0.119pt}} \multiput(816.00,538.17)(101.567,17.000){2}{\rule{1.309pt}{0.400pt}} \multiput(923.00,556.58)(2.737,0.491){17}{\rule{2.220pt}{0.118pt}} \multiput(923.00,555.17)(48.392,10.000){2}{\rule{1.110pt}{0.400pt}} \multiput(976.00,566.58)(2.263,0.492){21}{\rule{1.867pt}{0.119pt}} \multiput(976.00,565.17)(49.126,12.000){2}{\rule{0.933pt}{0.400pt}} \multiput(1029.00,578.58)(1.519,0.495){33}{\rule{1.300pt}{0.119pt}} \multiput(1029.00,577.17)(51.302,18.000){2}{\rule{0.650pt}{0.400pt}} \multiput(1083.00,596.58)(0.887,0.497){57}{\rule{0.807pt}{0.120pt}} \multiput(1083.00,595.17)(51.326,30.000){2}{\rule{0.403pt}{0.400pt}} \put(187,287){\usebox{\plotpoint}} \multiput(187.58,287.00)(0.498,0.919){81}{\rule{0.120pt}{0.833pt}} \multiput(186.17,287.00)(42.000,75.270){2}{\rule{0.400pt}{0.417pt}} \multiput(229.00,364.58)(0.731,0.498){71}{\rule{0.684pt}{0.120pt}} \multiput(229.00,363.17)(52.581,37.000){2}{\rule{0.342pt}{0.400pt}} \multiput(283.00,401.58)(1.299,0.498){79}{\rule{1.134pt}{0.120pt}} \multiput(283.00,400.17)(103.646,41.000){2}{\rule{0.567pt}{0.400pt}} \multiput(389.00,442.58)(2.000,0.497){51}{\rule{1.685pt}{0.120pt}} \multiput(389.00,441.17)(103.502,27.000){2}{\rule{0.843pt}{0.400pt}} \multiput(496.00,469.58)(2.714,0.496){37}{\rule{2.240pt}{0.119pt}} \multiput(496.00,468.17)(102.351,20.000){2}{\rule{1.120pt}{0.400pt}} \multiput(603.00,489.58)(3.174,0.495){31}{\rule{2.594pt}{0.119pt}} \multiput(603.00,488.17)(100.616,17.000){2}{\rule{1.297pt}{0.400pt}} \multiput(709.00,506.58)(3.410,0.494){29}{\rule{2.775pt}{0.119pt}} \multiput(709.00,505.17)(101.240,16.000){2}{\rule{1.388pt}{0.400pt}} \multiput(816.00,522.58)(3.204,0.495){31}{\rule{2.618pt}{0.119pt}} \multiput(816.00,521.17)(101.567,17.000){2}{\rule{1.309pt}{0.400pt}} \multiput(923.00,539.59)(3.465,0.488){13}{\rule{2.750pt}{0.117pt}} \multiput(923.00,538.17)(47.292,8.000){2}{\rule{1.375pt}{0.400pt}} \multiput(976.00,547.58)(2.477,0.492){19}{\rule{2.027pt}{0.118pt}} \multiput(976.00,546.17)(48.792,11.000){2}{\rule{1.014pt}{0.400pt}} \multiput(1029.00,558.58)(1.714,0.494){29}{\rule{1.450pt}{0.119pt}} \multiput(1029.00,557.17)(50.990,16.000){2}{\rule{0.725pt}{0.400pt}} \multiput(1083.00,574.58)(1.067,0.497){47}{\rule{0.948pt}{0.120pt}} \multiput(1083.00,573.17)(51.032,25.000){2}{\rule{0.474pt}{0.400pt}} \put(187,268){\usebox{\plotpoint}} \multiput(187.58,268.00)(0.498,0.919){81}{\rule{0.120pt}{0.833pt}} \multiput(186.17,268.00)(42.000,75.270){2}{\rule{0.400pt}{0.417pt}} \multiput(229.00,345.58)(0.731,0.498){71}{\rule{0.684pt}{0.120pt}} \multiput(229.00,344.17)(52.581,37.000){2}{\rule{0.342pt}{0.400pt}} \multiput(283.00,382.58)(1.267,0.498){81}{\rule{1.110pt}{0.120pt}} \multiput(283.00,381.17)(103.697,42.000){2}{\rule{0.555pt}{0.400pt}} \multiput(389.00,424.58)(2.078,0.497){49}{\rule{1.746pt}{0.120pt}} \multiput(389.00,423.17)(103.376,26.000){2}{\rule{0.873pt}{0.400pt}} \multiput(496.00,450.58)(2.714,0.496){37}{\rule{2.240pt}{0.119pt}} \multiput(496.00,449.17)(102.351,20.000){2}{\rule{1.120pt}{0.400pt}} \multiput(603.00,470.58)(2.994,0.495){33}{\rule{2.456pt}{0.119pt}} \multiput(603.00,469.17)(100.903,18.000){2}{\rule{1.228pt}{0.400pt}} \multiput(709.00,488.58)(3.644,0.494){27}{\rule{2.953pt}{0.119pt}} \multiput(709.00,487.17)(100.870,15.000){2}{\rule{1.477pt}{0.400pt}} \multiput(816.00,503.58)(3.204,0.495){31}{\rule{2.618pt}{0.119pt}} \multiput(816.00,502.17)(101.567,17.000){2}{\rule{1.309pt}{0.400pt}} \multiput(923.00,520.59)(3.058,0.489){15}{\rule{2.456pt}{0.118pt}} \multiput(923.00,519.17)(47.903,9.000){2}{\rule{1.228pt}{0.400pt}} \multiput(976.00,529.58)(2.737,0.491){17}{\rule{2.220pt}{0.118pt}} \multiput(976.00,528.17)(48.392,10.000){2}{\rule{1.110pt}{0.400pt}} \multiput(1029.00,539.58)(2.122,0.493){23}{\rule{1.762pt}{0.119pt}} \multiput(1029.00,538.17)(50.344,13.000){2}{\rule{0.881pt}{0.400pt}} \multiput(1083.00,552.58)(1.274,0.496){39}{\rule{1.110pt}{0.119pt}} \multiput(1083.00,551.17)(50.697,21.000){2}{\rule{0.555pt}{0.400pt}} \put(187,245){\usebox{\plotpoint}} \multiput(187.58,245.00)(0.498,0.907){81}{\rule{0.120pt}{0.824pt}} \multiput(186.17,245.00)(42.000,74.290){2}{\rule{0.400pt}{0.412pt}} \multiput(229.00,321.58)(0.731,0.498){71}{\rule{0.684pt}{0.120pt}} \multiput(229.00,320.17)(52.581,37.000){2}{\rule{0.342pt}{0.400pt}} \multiput(283.00,358.58)(1.267,0.498){81}{\rule{1.110pt}{0.120pt}} \multiput(283.00,357.17)(103.697,42.000){2}{\rule{0.555pt}{0.400pt}} \multiput(389.00,400.58)(2.078,0.497){49}{\rule{1.746pt}{0.120pt}} \multiput(389.00,399.17)(103.376,26.000){2}{\rule{0.873pt}{0.400pt}} \multiput(496.00,426.58)(2.714,0.496){37}{\rule{2.240pt}{0.119pt}} \multiput(496.00,425.17)(102.351,20.000){2}{\rule{1.120pt}{0.400pt}} \multiput(603.00,446.58)(2.994,0.495){33}{\rule{2.456pt}{0.119pt}} \multiput(603.00,445.17)(100.903,18.000){2}{\rule{1.228pt}{0.400pt}} \multiput(709.00,464.58)(3.410,0.494){29}{\rule{2.775pt}{0.119pt}} \multiput(709.00,463.17)(101.240,16.000){2}{\rule{1.388pt}{0.400pt}} \multiput(816.00,480.58)(3.204,0.495){31}{\rule{2.618pt}{0.119pt}} \multiput(816.00,479.17)(101.567,17.000){2}{\rule{1.309pt}{0.400pt}} \multiput(923.00,497.59)(3.465,0.488){13}{\rule{2.750pt}{0.117pt}} \multiput(923.00,496.17)(47.292,8.000){2}{\rule{1.375pt}{0.400pt}} \multiput(976.00,505.58)(2.737,0.491){17}{\rule{2.220pt}{0.118pt}} \multiput(976.00,504.17)(48.392,10.000){2}{\rule{1.110pt}{0.400pt}} \multiput(1029.00,515.58)(2.122,0.493){23}{\rule{1.762pt}{0.119pt}} \multiput(1029.00,514.17)(50.344,13.000){2}{\rule{0.881pt}{0.400pt}} \multiput(1083.00,528.58)(1.580,0.495){31}{\rule{1.347pt}{0.119pt}} \multiput(1083.00,527.17)(50.204,17.000){2}{\rule{0.674pt}{0.400pt}} \put(187,213){\usebox{\plotpoint}} \multiput(187.58,213.00)(0.498,0.907){81}{\rule{0.120pt}{0.824pt}} \multiput(186.17,213.00)(42.000,74.290){2}{\rule{0.400pt}{0.412pt}} \multiput(229.00,289.58)(0.751,0.498){69}{\rule{0.700pt}{0.120pt}} \multiput(229.00,288.17)(52.547,36.000){2}{\rule{0.350pt}{0.400pt}} \multiput(283.00,325.58)(1.267,0.498){81}{\rule{1.110pt}{0.120pt}} \multiput(283.00,324.17)(103.697,42.000){2}{\rule{0.555pt}{0.400pt}} \multiput(389.00,367.58)(2.078,0.497){49}{\rule{1.746pt}{0.120pt}} \multiput(389.00,366.17)(103.376,26.000){2}{\rule{0.873pt}{0.400pt}} \multiput(496.00,393.58)(2.714,0.496){37}{\rule{2.240pt}{0.119pt}} \multiput(496.00,392.17)(102.351,20.000){2}{\rule{1.120pt}{0.400pt}} \multiput(603.00,413.58)(2.833,0.495){35}{\rule{2.332pt}{0.119pt}} \multiput(603.00,412.17)(101.161,19.000){2}{\rule{1.166pt}{0.400pt}} \multiput(709.00,432.58)(3.410,0.494){29}{\rule{2.775pt}{0.119pt}} \multiput(709.00,431.17)(101.240,16.000){2}{\rule{1.388pt}{0.400pt}} \multiput(816.00,448.58)(3.204,0.495){31}{\rule{2.618pt}{0.119pt}} \multiput(816.00,447.17)(101.567,17.000){2}{\rule{1.309pt}{0.400pt}} \multiput(923.00,465.59)(3.058,0.489){15}{\rule{2.456pt}{0.118pt}} \multiput(923.00,464.17)(47.903,9.000){2}{\rule{1.228pt}{0.400pt}} \multiput(976.00,474.59)(3.058,0.489){15}{\rule{2.456pt}{0.118pt}} \multiput(976.00,473.17)(47.903,9.000){2}{\rule{1.228pt}{0.400pt}} \multiput(1029.00,483.58)(2.122,0.493){23}{\rule{1.762pt}{0.119pt}} \multiput(1029.00,482.17)(50.344,13.000){2}{\rule{0.881pt}{0.400pt}} \multiput(1083.00,496.58)(1.797,0.494){27}{\rule{1.513pt}{0.119pt}} \multiput(1083.00,495.17)(49.859,15.000){2}{\rule{0.757pt}{0.400pt}} \put(187,159){\usebox{\plotpoint}} \multiput(187.58,159.00)(0.498,0.931){81}{\rule{0.120pt}{0.843pt}} \multiput(186.17,159.00)(42.000,76.251){2}{\rule{0.400pt}{0.421pt}} \multiput(229.00,237.58)(0.773,0.498){67}{\rule{0.717pt}{0.120pt}} \multiput(229.00,236.17)(52.512,35.000){2}{\rule{0.359pt}{0.400pt}} \multiput(283.00,272.58)(1.299,0.498){79}{\rule{1.134pt}{0.120pt}} \multiput(283.00,271.17)(103.646,41.000){2}{\rule{0.567pt}{0.400pt}} \multiput(389.00,313.58)(2.078,0.497){49}{\rule{1.746pt}{0.120pt}} \multiput(389.00,312.17)(103.376,26.000){2}{\rule{0.873pt}{0.400pt}} \multiput(496.00,339.58)(2.714,0.496){37}{\rule{2.240pt}{0.119pt}} \multiput(496.00,338.17)(102.351,20.000){2}{\rule{1.120pt}{0.400pt}} \multiput(603.00,359.58)(2.833,0.495){35}{\rule{2.332pt}{0.119pt}} \multiput(603.00,358.17)(101.161,19.000){2}{\rule{1.166pt}{0.400pt}} \multiput(709.00,378.58)(3.410,0.494){29}{\rule{2.775pt}{0.119pt}} \multiput(709.00,377.17)(101.240,16.000){2}{\rule{1.388pt}{0.400pt}} \multiput(816.00,394.58)(3.022,0.495){33}{\rule{2.478pt}{0.119pt}} \multiput(816.00,393.17)(101.857,18.000){2}{\rule{1.239pt}{0.400pt}} \multiput(923.00,412.59)(3.058,0.489){15}{\rule{2.456pt}{0.118pt}} \multiput(923.00,411.17)(47.903,9.000){2}{\rule{1.228pt}{0.400pt}} \multiput(976.00,421.58)(2.737,0.491){17}{\rule{2.220pt}{0.118pt}} \multiput(976.00,420.17)(48.392,10.000){2}{\rule{1.110pt}{0.400pt}} \multiput(1029.00,431.58)(2.122,0.493){23}{\rule{1.762pt}{0.119pt}} \multiput(1029.00,430.17)(50.344,13.000){2}{\rule{0.881pt}{0.400pt}} \multiput(1083.00,444.58)(1.929,0.494){25}{\rule{1.614pt}{0.119pt}} \multiput(1083.00,443.17)(49.649,14.000){2}{\rule{0.807pt}{0.400pt}} \end{picture} {\tenrm\baselineskip=12pt \noindent Figure 2: The ratio $\delta(x_l,y_l)$ (\ref{delxy}) vs. $y_l$ for various $x_l$ values. Top curve is for $x_l = 0.01$, the next for $x_l=0.1$. Each subsequent curve represents an increase of $x_l$ by $0.1$.} \vglue 0.2cm For comparison with the ${\mathrm{e}}$p case we now show in Fig.2 the correction $\delta(x_l,y_l)$, using (\ref{nsigBorn}), (\ref{bremscross}) and (\ref{Compton}), defined by \begin{equation} \frac{{\mathrm{d}}^2\sigma^{\mathrm{Br}}}{{\mathrm{d}} x_l {\mathrm{d}} y_l} + \frac{{\mathrm{d}}^2\sigma^{\mathrm{C}}}{{\mathrm{d}} x_l {\mathrm{d}} y_l}= \frac{{\mathrm{d}}^2\sigma^{\mathrm{B}}}{{\mathrm{d}} x_l {\mathrm{d}} y_l}\,\,\delta(x_l,y_l)\ . \label{delxy} \end{equation} Note that we use here the leptonic variables to conform with the ${\mathrm{e}}$p case. A closer examination reveals that initial and final state radiation are similar in order of magnitude throughout most of the $x,y$ region, whereas the Compton contributions is appreciable only in the small and medium $x$, large $y$ region. Also we note that if one freezes $F_2$ at $F_2(Q_0^2)$ for $Q^2<Q^2_0$, instead of extrapolating as in (\ref{extrapol}), we find that only the $x_l=0.01$ curve changes significantly. It decreases at small and medium $y_l$ by up to 50\%. Note that inclusion of final state radiation implies a perfect measurement of the energy and momentum of the tagged lepton, even in the presence of collinear radiation. We see in Fig.2 that radiative effects can in principle be large, around 40\% for medium $x$ and small $y$. Next we show in Fig.3 the correction factor $\delta(x_h,Q^2_h)$, defined by \begin{equation} \frac{{\mathrm{d}}^2\sigma^{\mathrm{corr}}}{{\mathrm{d}} x_h {\mathrm{d}} Q^2_h} = \frac{{\mathrm{d}}^2\sigma^{\mathrm{B}}}{{\mathrm{d}} x_h {\mathrm{d}} Q^2_h} \delta(x_h,Q^2_h)\ . \label{delxQ2}, \end{equation} in terms of hadronic variables, as a function of $x_h$ for various choices of $Q_h^2$, cf. (\ref{hadcorr}). Here only initial state lepton bremsstrahlung is taken into account, because the scattered lepton is not used in constructing the kinematic variables. \vglue 0.5cm \setlength{\unitlength}{0.240900pt} \ifx\plotpoint\undefined\newsavebox{\plotpoint}\fi \sbox{\plotpoint}{\rule[-0.200pt]{0.400pt}{0.400pt}}% \begin{picture}(1500,900)(0,0) \font\gnuplot=cmr10 at 10pt \gnuplot \sbox{\plotpoint}{\rule[-0.200pt]{0.400pt}{0.400pt}}% \put(176.0,113.0){\rule[-0.200pt]{0.400pt}{184.048pt}} \put(176.0,113.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,113){\makebox(0,0)[r]{-0.4}} \put(1416.0,113.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,209.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,209){\makebox(0,0)[r]{-0.35}} \put(1416.0,209.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,304.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,304){\makebox(0,0)[r]{-0.3}} \put(1416.0,304.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,399.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,399){\makebox(0,0)[r]{-0.25}} \put(1416.0,399.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,495.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,495){\makebox(0,0)[r]{-0.2}} \put(1416.0,495.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,590.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,590){\makebox(0,0)[r]{-0.15}} \put(1416.0,590.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,686.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,686){\makebox(0,0)[r]{-0.1}} \put(1416.0,686.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,782.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,782){\makebox(0,0)[r]{-0.05}} \put(1416.0,782.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,877.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(154,877){\makebox(0,0)[r]{0}} \put(1416.0,877.0){\rule[-0.200pt]{4.818pt}{0.400pt}} \put(176.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(176,68){\makebox(0,0){0}} \put(176.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(302.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(302,68){\makebox(0,0){0.1}} \put(302.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(428.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(428,68){\makebox(0,0){0.2}} \put(428.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(554.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(554,68){\makebox(0,0){0.3}} \put(554.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(680.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(680,68){\makebox(0,0){0.4}} \put(680.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(806.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(806,68){\makebox(0,0){0.5}} \put(806.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(932.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(932,68){\makebox(0,0){0.6}} \put(932.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1058.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1058,68){\makebox(0,0){0.7}} \put(1058.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1184.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1184,68){\makebox(0,0){0.8}} \put(1184.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1310.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1310,68){\makebox(0,0){0.9}} \put(1310.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1436.0,113.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(1436,68){\makebox(0,0){1}} \put(1436.0,857.0){\rule[-0.200pt]{0.400pt}{4.818pt}} \put(176.0,113.0){\rule[-0.200pt]{303.534pt}{0.400pt}} \put(1436.0,113.0){\rule[-0.200pt]{0.400pt}{184.048pt}} \put(176.0,877.0){\rule[-0.200pt]{303.534pt}{0.400pt}} \put(806,23){\makebox(0,0){$x_h$}} \put(176.0,113.0){\rule[-0.200pt]{0.400pt}{184.048pt}} \put(189,698){\usebox{\plotpoint}} \multiput(189.00,698.58)(0.625,0.498){77}{\rule{0.600pt}{0.120pt}} \multiput(189.00,697.17)(48.755,40.000){2}{\rule{0.300pt}{0.400pt}} \multiput(239.00,738.58)(2.479,0.493){23}{\rule{2.038pt}{0.119pt}} \multiput(239.00,737.17)(58.769,13.000){2}{\rule{1.019pt}{0.400pt}} \multiput(302.00,751.59)(6.944,0.477){7}{\rule{5.140pt}{0.115pt}} \multiput(302.00,750.17)(52.332,5.000){2}{\rule{2.570pt}{0.400pt}} \multiput(365.00,756.60)(9.108,0.468){5}{\rule{6.400pt}{0.113pt}} \multiput(365.00,755.17)(49.716,4.000){2}{\rule{3.200pt}{0.400pt}} \put(428,760.17){\rule{12.700pt}{0.400pt}} \multiput(428.00,759.17)(36.641,2.000){2}{\rule{6.350pt}{0.400pt}} \put(491,762.17){\rule{12.700pt}{0.400pt}} \multiput(491.00,761.17)(36.641,2.000){2}{\rule{6.350pt}{0.400pt}} \multiput(554.00,764.61)(13.858,0.447){3}{\rule{8.500pt}{0.108pt}} \multiput(554.00,763.17)(45.358,3.000){2}{\rule{4.250pt}{0.400pt}} \put(617,765.67){\rule{15.177pt}{0.400pt}} \multiput(617.00,766.17)(31.500,-1.000){2}{\rule{7.588pt}{0.400pt}} \multiput(680.00,766.59)(6.944,0.477){7}{\rule{5.140pt}{0.115pt}} \multiput(680.00,765.17)(52.332,5.000){2}{\rule{2.570pt}{0.400pt}} \put(743,771.17){\rule{12.700pt}{0.400pt}} \multiput(743.00,770.17)(36.641,2.000){2}{\rule{6.350pt}{0.400pt}} \put(806,772.67){\rule{15.177pt}{0.400pt}} \multiput(806.00,772.17)(31.500,1.000){2}{\rule{7.588pt}{0.400pt}} \put(932,773.67){\rule{15.177pt}{0.400pt}} \multiput(932.00,773.17)(31.500,1.000){2}{\rule{7.588pt}{0.400pt}} \multiput(995.00,773.95)(13.858,-0.447){3}{\rule{8.500pt}{0.108pt}} \multiput(995.00,774.17)(45.358,-3.000){2}{\rule{4.250pt}{0.400pt}} \multiput(1058.00,772.61)(13.858,0.447){3}{\rule{8.500pt}{0.108pt}} \multiput(1058.00,771.17)(45.358,3.000){2}{\rule{4.250pt}{0.400pt}} \put(1121,775.17){\rule{12.700pt}{0.400pt}} \multiput(1121.00,774.17)(36.641,2.000){2}{\rule{6.350pt}{0.400pt}} \multiput(1184.00,775.95)(13.858,-0.447){3}{\rule{8.500pt}{0.108pt}} \multiput(1184.00,776.17)(45.358,-3.000){2}{\rule{4.250pt}{0.400pt}} \multiput(1247.00,774.61)(13.858,0.447){3}{\rule{8.500pt}{0.108pt}} \multiput(1247.00,773.17)(45.358,3.000){2}{\rule{4.250pt}{0.400pt}} \put(1310,777.17){\rule{12.700pt}{0.400pt}} \multiput(1310.00,776.17)(36.641,2.000){2}{\rule{6.350pt}{0.400pt}} \put(869.0,774.0){\rule[-0.200pt]{15.177pt}{0.400pt}} \put(189,543){\usebox{\plotpoint}} \multiput(189.58,543.00)(0.498,0.993){97}{\rule{0.120pt}{0.892pt}} \multiput(188.17,543.00)(50.000,97.149){2}{\rule{0.400pt}{0.446pt}} \multiput(239.00,642.58)(1.132,0.497){53}{\rule{1.000pt}{0.120pt}} \multiput(239.00,641.17)(60.924,28.000){2}{\rule{0.500pt}{0.400pt}} \multiput(302.00,670.58)(2.139,0.494){27}{\rule{1.780pt}{0.119pt}} \multiput(302.00,669.17)(59.306,15.000){2}{\rule{0.890pt}{0.400pt}} \multiput(365.00,685.59)(3.640,0.489){15}{\rule{2.900pt}{0.118pt}} \multiput(365.00,684.17)(56.981,9.000){2}{\rule{1.450pt}{0.400pt}} \multiput(428.00,694.59)(6.944,0.477){7}{\rule{5.140pt}{0.115pt}} \multiput(428.00,693.17)(52.332,5.000){2}{\rule{2.570pt}{0.400pt}} \multiput(491.00,699.59)(5.644,0.482){9}{\rule{4.300pt}{0.116pt}} \multiput(491.00,698.17)(54.075,6.000){2}{\rule{2.150pt}{0.400pt}} \multiput(554.00,705.59)(5.644,0.482){9}{\rule{4.300pt}{0.116pt}} \multiput(554.00,704.17)(54.075,6.000){2}{\rule{2.150pt}{0.400pt}} \multiput(680.00,711.60)(9.108,0.468){5}{\rule{6.400pt}{0.113pt}} \multiput(680.00,710.17)(49.716,4.000){2}{\rule{3.200pt}{0.400pt}} \put(617.0,711.0){\rule[-0.200pt]{15.177pt}{0.400pt}} \multiput(806.00,715.60)(9.108,0.468){5}{\rule{6.400pt}{0.113pt}} \multiput(806.00,714.17)(49.716,4.000){2}{\rule{3.200pt}{0.400pt}} \multiput(869.00,719.61)(13.858,0.447){3}{\rule{8.500pt}{0.108pt}} \multiput(869.00,718.17)(45.358,3.000){2}{\rule{4.250pt}{0.400pt}} \put(932,722.17){\rule{12.700pt}{0.400pt}} \multiput(932.00,721.17)(36.641,2.000){2}{\rule{6.350pt}{0.400pt}} \put(995,722.67){\rule{15.177pt}{0.400pt}} \multiput(995.00,723.17)(31.500,-1.000){2}{\rule{7.588pt}{0.400pt}} \multiput(1058.00,723.61)(13.858,0.447){3}{\rule{8.500pt}{0.108pt}} \multiput(1058.00,722.17)(45.358,3.000){2}{\rule{4.250pt}{0.400pt}} \put(1121,725.67){\rule{15.177pt}{0.400pt}} \multiput(1121.00,725.17)(31.500,1.000){2}{\rule{7.588pt}{0.400pt}} \put(743.0,715.0){\rule[-0.200pt]{15.177pt}{0.400pt}} \multiput(1247.00,727.61)(13.858,0.447){3}{\rule{8.500pt}{0.108pt}} \multiput(1247.00,726.17)(45.358,3.000){2}{\rule{4.250pt}{0.400pt}} \put(1184.0,727.0){\rule[-0.200pt]{15.177pt}{0.400pt}} \multiput(194.58,113.00)(0.498,3.467){87}{\rule{0.120pt}{2.856pt}} \multiput(193.17,113.00)(45.000,304.073){2}{\rule{0.400pt}{1.428pt}} \multiput(239.58,423.00)(0.499,0.627){123}{\rule{0.120pt}{0.602pt}} \multiput(238.17,423.00)(63.000,77.751){2}{\rule{0.400pt}{0.301pt}} \multiput(302.00,502.58)(0.930,0.498){65}{\rule{0.841pt}{0.120pt}} \multiput(302.00,501.17)(61.254,34.000){2}{\rule{0.421pt}{0.400pt}} \multiput(365.00,536.58)(1.593,0.496){37}{\rule{1.360pt}{0.119pt}} \multiput(365.00,535.17)(60.177,20.000){2}{\rule{0.680pt}{0.400pt}} \multiput(428.00,556.58)(1.881,0.495){31}{\rule{1.582pt}{0.119pt}} \multiput(428.00,555.17)(59.716,17.000){2}{\rule{0.791pt}{0.400pt}} \multiput(491.00,573.58)(2.693,0.492){21}{\rule{2.200pt}{0.119pt}} \multiput(491.00,572.17)(58.434,12.000){2}{\rule{1.100pt}{0.400pt}} \multiput(554.00,585.59)(4.126,0.488){13}{\rule{3.250pt}{0.117pt}} \multiput(554.00,584.17)(56.254,8.000){2}{\rule{1.625pt}{0.400pt}} \multiput(617.00,593.59)(3.640,0.489){15}{\rule{2.900pt}{0.118pt}} \multiput(617.00,592.17)(56.981,9.000){2}{\rule{1.450pt}{0.400pt}} \multiput(680.00,602.59)(5.644,0.482){9}{\rule{4.300pt}{0.116pt}} \multiput(680.00,601.17)(54.075,6.000){2}{\rule{2.150pt}{0.400pt}} \multiput(743.00,608.59)(6.944,0.477){7}{\rule{5.140pt}{0.115pt}} \multiput(743.00,607.17)(52.332,5.000){2}{\rule{2.570pt}{0.400pt}} \multiput(806.00,613.59)(6.944,0.477){7}{\rule{5.140pt}{0.115pt}} \multiput(806.00,612.17)(52.332,5.000){2}{\rule{2.570pt}{0.400pt}} \multiput(869.00,618.60)(9.108,0.468){5}{\rule{6.400pt}{0.113pt}} \multiput(869.00,617.17)(49.716,4.000){2}{\rule{3.200pt}{0.400pt}} \multiput(932.00,622.59)(5.644,0.482){9}{\rule{4.300pt}{0.116pt}} \multiput(932.00,621.17)(54.075,6.000){2}{\rule{2.150pt}{0.400pt}} \put(995,627.67){\rule{15.177pt}{0.400pt}} \multiput(995.00,627.17)(31.500,1.000){2}{\rule{7.588pt}{0.400pt}} \put(1058,629.17){\rule{12.700pt}{0.400pt}} \multiput(1058.00,628.17)(36.641,2.000){2}{\rule{6.350pt}{0.400pt}} \multiput(1121.00,631.59)(6.944,0.477){7}{\rule{5.140pt}{0.115pt}} \multiput(1121.00,630.17)(52.332,5.000){2}{\rule{2.570pt}{0.400pt}} \multiput(1184.00,636.61)(13.858,0.447){3}{\rule{8.500pt}{0.108pt}} \multiput(1184.00,635.17)(45.358,3.000){2}{\rule{4.250pt}{0.400pt}} \put(1247,638.67){\rule{15.177pt}{0.400pt}} \multiput(1247.00,638.17)(31.500,1.000){2}{\rule{7.588pt}{0.400pt}} \put(1310,640.17){\rule{12.700pt}{0.400pt}} \multiput(1310.00,639.17)(36.641,2.000){2}{\rule{6.350pt}{0.400pt}} \put(1310.0,730.0){\rule[-0.200pt]{15.177pt}{0.400pt}} \end{picture} {\tenrm\baselineskip=12pt \noindent Figure 3: The ratio $\delta(x_h,Q^2_h)$ vs. $x_h$ for three $Q^2_h$ values. Top curve: $Q^2_h = 1 \, {\rm GeV^2}$. Middle curve: $Q^2_h = 10\, {\rm GeV^2}$. Lower curve: $Q^2_h = 100\, {\rm GeV^2}$.} \vglue 0.5cm We see that the corrections are sizable only for large $Q^2$. Using however the resummed version of (\ref{resummed2}) and (\ref{resummedL}), we find that the corrections are reduced by about an order of magnitude. Thus we conclude that the size of the radiative corrections depends significantly on the set of variables chosen. These corrections can in principle be quite large. As noted before, in practice mixed variables are used. In view of the results obtained in this paper, we think a more careful study, involving Monte Carlo simulation of the full final state, is warranted.
\section{1. Introduction} In this paper, we study a general class of two-dimensional higher-derivative gravitational theories in a Riemann-Cartan spacetime. As is well known, in two dimensions the Einstein-Hilbert action is a total derivative and hence cannot be used to construct a two-dimensional version of general relativity. It is however possible to construct actions whose lagrangian density is given by an arbitrary power of the Ricci scalar [1]. These models turn out to be equivalent to the ordinary Einstein-Hilbert action non-minimally coupled to a scalar field with a power-law potential [2-4]. In such formalism the field equations become second order and it is possible to obtain exact solutions and to perform the Dirac quantization of the theory [5]. Some well-known special cases of gravity-scalar theories in two dimensions include the Jackiw-Teitelboim [6] and the "string" models [7-8]. Another useful generalization of two-dimensional gravity is given by the consideration of \RC geometries with non-trivial torsion. Several authors have studied various aspects of a model with action quadratic in the curvature and the torsion [9-15]. It has been proved that this model is completely integrable [9-10] and, exploiting the hamiltonian formalism, it has been shown that its symmetries are a realization of a specific non-linear algebra [11-12] and that its quantization can be performed exactly [13]. Finally, a further interesting aspect of some two-dimensional models is that they can be interpreted as gauge theories of the \poi or de Sitter group in two dimensions [16]. In ref. [14] it has been shown that this interpretation can be extended to theories with non-trivial torsion, provided that one generalizes the notion of gauge invariance to groups generated by non-linear algebras. In this paper we try to extend the results obtained so far to the case of actions containing arbitrary powers of the curvature and quadratic torsion. We show that the static solutions of the field equations can be found exactly and discuss their geometry. We also investigate the hamiltonian formulation of the models and find the constraint algebra, which is a non-linear deformation of the two-dimensional \poi algebra and therefore permits their interpretation as a non-linear gauge theory of such algebra. Finally, we perform the Dirac quantization of the model and define the space of the physical states. \section{2. The action and the field equations} We consider a 2-dimensional lorentzian manifold with signature $(-,+)$ endowed with a \RC geometry. The geometry can be described by the zweibein field $e^a_\m$ and the Lorentz connection $\conn{\m}$, where $a,b,..$ are tangent space indices which can take the values $0$, $1$ and $\m,\n,...$ are world indices, whose values will be denoted by $t$, $x$. The curvature and torsion are defined as $$\eqalign{&R\ba_\mn=\de_\m\conn{\n}-\de_\n\conn{\m}\cr &T^a_\mn=\de_\m e^a_\n-\de_\n e^a_\m+\o{^a_{\ b}}}\def\bam{{_a^{\ b}}}\def\ba{^{ab}_\m e^b_\n-\o{^a_{\ b}}}\def\bam{{_a^{\ b}}}\def\ba{^{ab}_\n e^b_\m\cr} \eqno(1)$$ and the Ricci scalar as $R=e_a^\m e_b^\n R\ba_\mn$. In two dimensions the Ricci scalar determines uniquely the Riemann tensor by the relation $\curv abcd=-\ha\e_{ab}\e_{cd}R$. Moreover, the connection can be written as $$\conn{\m}=\e\ba\o_\m$$ where $\e\ba$ is the antisymmetric tensor $\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r}=-\e\ba$, $\e_{01}=1$. In this paper, we study higher derivative actions of the form: $$S=\edx (R^k-\gh T^2)\eqno(2)$$ where $e=$det $e^a_\m$, $T^2=\tors abc T^{abc}$. Here, $k$ is any real number except $0$, $1$ and $\g$ a coupling constant. These actions generalize to \RC geometry the higher-derivative actions introduced in [1] and further investigated in [2-5]. The special case $k=2$ has already been studied by several authors [9-15]. By defining a scalar field $\y=kR^{k-1}$, it is possible, by a standard argument [2-4] to reduce the action (2) to a form which is linear in the curvature: $$S'=\edx\left(\y R-\gh T^2+\L\y^h\right)\eqno(3)$$ where $\L=(1-k)k^{-k/(k-1)}$ and $h={k\over k-1}$. The action can be further reduced to a fully first order form, by introducing a doublet of scalar fields $\y_a$ [13]: $$S''=\edx\left(\y R+\y_a\, ^*T^a-\gi\y_a\y^a+\L\y^h\right)\eqno(4)$$ where $^*T^a=\e^{bc}\tors abc$. The field equations obtained by varying (3) with respect to the fields $\y$, $e^a$ and $\o$ can be written as: $$\eqalign{&R+h\L\y^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g}=0\cr &-\na^b\tors cdb +T\ba_{\ \ c}\tors abd-{1\over 4}}\def\di{{\rm d}}\def\pro{\propto}\def\app{\approx g_{cd}(T^2-{2\L\over\g}\y^h)=0\cr &\e^a_{\ c}e_a^\m\de_\m\y-\g\e\ba\tors abc=0\cr}\eqno(5)$$ \section{3. The static solutions} In the following, we shall look for the static solutions of these equations. According to the discussion of the special case $k=2$ performed in ref. [9], it results convenient to seek for the solutions in a conformal gauge. We therefore adopt the ansatz: $$e^0_t=e^1_x=e^{2\r(x)}\qquad\qquad e^0_x=e^1_t=0$$ We also assume that $\o_\m=\ep\de^\n\c$, with $\c=\c(x)$, which yields $$\o_x=0,\qquad\qquad \o_t=\c'$$ where a prime denotes derivative with respect to $x$. In terms of these variables, one has: $$R=2e^{-2\r}\c'',\qquad\qquad\tors 001=e^{-\r}(\r'-\c')\eqno(6)$$ and the field equations become: $$\c''+h{\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x\y^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g}\erp=0\eqno(7.a)$$ $$\y'-\g(\r'-\c')=0\eqno(7.b)$$ $$\r''-\c''-\ha(\r'-\c')(\r'+\c')+{\L\over 2\g}\y^h\erp=0\eqno(7.c)$$ $$\ha(\r'-\c')(\r'+\c')+{\L\over 2\g}\y^h\erp=0\eqno(7.d)$$ These equations admit a special solution $\c=\r=\y=0$, corresponding to a manifold with vanishing curvature and torsion (we are cosidering the case of vanishing cosmological constant). The zero torsion solutions do not therefore reduce to the extremals of the action (3) with $\g=0$, which have been discussed in [5]. This fact has been observed in [9] for the special case $k=2$. One can however obtain more general solutions to (7). Combining (7.c) and (7.d) one gets $$\r''-\c''=(\r'-\c')(\r'+\c')=-{\L\over\g}\y^h\erp\eqno(8)$$ A first integral of the first equation (8) is $$\r=\ha(f+\ln(Ef'))\eqno(9)$$ where $f\equiv\r-\c$ and $E$ is an integration constant. {}From (7.b), one has $\y=\g f$ and thus (8) yields $$f''=\L E\g^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g} e^ff^hf'\eqno(10)$$ which can be integrated to give $$f'=\L E\g^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g}\int^f_0 g^he^gdg\ +A\eqno(11)$$ where the integral on the r.h.s. is proportional to the incomplete gamma function $\G(h+1,-f)$ and $A$ is an integration constant. The equation (11) can then be integrated numerically to give $f$. This solution generalizes that obtained in ref. [9] for $h=2$. One can now express the curvature and the torsion in terms of the function $f$: $$R=-\L(\g f)^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g}\qquad\qquad T^2=e^{-f}f'\eqno(12)$$ It is then easy to study the qualitative behaviour of the solutions of (11). We shall impose the positivity of $f$ in order to avoid problems when $h$ is not integer. From the asymptotics of (11), follows that $f\to\infty}\def\id{\equiv}\def\mo{{-1}}\def\ha{{1\over 2}$ for a finite value of $x$. If $h>1$, a curvature singularity is therefore always present at finite $x$ in these coordinates, while the scalar $T^2$ is singular if $h>0$ as $f\to\infty}\def\id{\equiv}\def\mo{{-1}}\def\ha{{1\over 2}$. A numerical study permits to distinguish three different possible behaviours for $f$ (see fig. 1-3): a) If $h>-1$ and $A>0$, $f$ grows monotonically from $0$ to $\infty}\def\id{\equiv}\def\mo{{-1}}\def\ha{{1\over 2}$ between two finite values \xt\ and $x_1$}\def\xt{$x_2$}\def\conn#1{\o\ba_{\ \ #1}\ of $x$. b) If $h>-1$ and $A<0$, the solution has two branches: one of them decreases monotonically between a constant value $f_0$ at $x=-\infty}\def\id{\equiv}\def\mo{{-1}}\def\ha{{1\over 2}$ and 0 at $x=$ $x_1$}\def\xt{$x_2$}\def\conn#1{\o\ba_{\ \ #1}, while the other grows monotonically between $f_0$ at $-\infty}\def\id{\equiv}\def\mo{{-1}}\def\ha{{1\over 2}$ and infinity at $x=$\xt. c) If $h<-1$, for any $A$, the behaviour of $f$ is similar to the case b), but $f'(x_2)\to-\infty}\def\id{\equiv}\def\mo{{-1}}\def\ha{{1\over 2}$. In order to investigate the properties of the solutions near the critical point, one must study the behavior of the functions $R$, $T^2$ and $\erp$ near $x_1$}\def\xt{$x_2$}\def\conn#1{\o\ba_{\ \ #1}\ and \xt. It is easy to check that for $f\to\infty}\def\id{\equiv}\def\mo{{-1}}\def\ha{{1\over 2}$, $R\sim f^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g}$, $T^2\sim f^h$, $\erp\sim e^{2f}$, while for $f\to 0$, $R\sim f^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g}$, $T^2\sim\const +f^{h+1}$, $\erp\sim\const +f^{h+1}$. Depending on the value of $h$, one can then distinguish several cases which are summarized in the tables 1-3. The following general results can be stated: If $h>-1$ and $A>0$, a naked singularity is always present, either at $x_1$}\def\xt{$x_2$}\def\conn#1{\o\ba_{\ \ #1}\ or at \xt. More interesting are the cases where $h>-1$ and $A<0$ or $h>-1$. In these cases, the two branches of the solution describe the interior and the exterior of an asymptotically flat }\def\hd{higher derivative }\def\st{spacetime \bh with the horizon located at $x=-\infty}\def\id{\equiv}\def\mo{{-1}}\def\ha{{1\over 2}$. For $h<1$ (resp. $h>1$), the curvature singularity is at $x_1$}\def\xt{$x_2$}\def\conn#1{\o\ba_{\ \ #1} (resp. \xt), while spatial infinity is at \xt\ (resp. $x_1$}\def\xt{$x_2$}\def\conn#1{\o\ba_{\ \ #1}). For $0<h<1$, however, the torsion diverges at spatial infinity. Of course, a detailed study of the spacetime structure would require a more explicit form of the solutions. For the special cases $h=0$ and $h=1$, this will be afforded in a future paper. \section{4. First order formalism} In terms of differential forms, $e^a=e^a_\m dx^\m$, $\o=\o_\m dx^\m$, the first order lagrangian in (4) can be written as: $$\ha{\cal L}}\def\de{\partial}\def\na{\nabla}\def\per{\times = \y_2 d\o+\y_aT^a+\left(\Lq\y_2^h+{1\over 2\g}\y_c\y^c\right)\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r} e^ae^b \eqno(13)$$ where we have renamed $\y$ as $\y_2$. This form of the lagrangian is especially convenient because it permits to evidentiate the connection of our models with the formulation of 2-dimensional gravity as a gauge theory of the Poincar\'e group ISO(1,1) or one of its generalizations [16,8]. In this formalism, $e^a$ and $\o$ play the role of gauge connections, while the $\y$ are considered as auxiliary fields. Local Poincar\'e transformations }\def\ther{thermodynamical }\def\coo{coordinates with parameters $\x^2$ and $\x^a$, corresponding to Lorentz rotations and to translations, act infinitesimally on the fields according to: $$\d e^a=d\x^a+\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r}(\x^b\o-\x^2e^b)\qquad\qquad\d\o=d\x^2$$ $$\d\y_a=\e\bam\x^2\y_b\qquad\qquad\d\y_2=\e\bam\x^a\y_b$$ and $R=d\o$ and $T^a$ are the field strengths corresponding to Lorentz rotations and translations respectively. The first two terms in the lagrangian (13) are invariant under these transformations }\def\ther{thermodynamical }\def\coo{coordinates , while the potential terms are not. As we shall see, the full action is in fact invariant under a non-linear generalization of the \poi group. In first order formalism, the field equations are given by: $$\eqalign{&T^a+\gi\y^a\e_{bc}e^be^c=0\cr &d\o+\Lq\y_2^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g}\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r} e^ae^b=0\cr &d\y_a+\y_b\e\bam\o+\left({\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x\y_2^h+\gi\y_c\y^c\right)\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r} e^b=0\cr &d\y_2+\y_a\e{^a_{\ b}}}\def\bam{{_a^{\ b}}}\def\ba{^{ab}\e^b=0\cr}\eqno(14)$$ whose static solutions can be written in terms of the function $f$ defined above as: $$\e^0_t=e^1_x=\sqrt{Ef'e^f}\qquad\qquad$$ $$\o_t={\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x E\g^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g} f^he^f-\ha f'\qquad\qquad\o_x=0$$ $$\y_0=\g\sqrt{f'\over Ee^f}\qquad\qquad\y_1=0\qquad\qquad\y_2=\g f$$ \section{5. The hamiltonian formalism} Another advantage of the first order formalism is that it leads naturally to a hamiltonian formulation of the model, and hence permits a straightforward discussion of its symmetries and quantization. In fact, after integration by parts, the lagrangian density can be written as: $$\eqalign{\ha{\cal L}}\def\de{\partial}\def\na{\nabla}\def\per{\times=&\y_a\dot e^a_x+\y_2\dot\o_x\cr +&e^a_t(\y_a'+\e\bam\y_b\o_x+{\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x\y_2^h\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r} e_x^b+\gi \y_c\y^c\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r} e_x^b) +\o_t(\y_2'+\y_a\e{^a_{\ b}}}\def\bam{{_a^{\ b}}}\def\ba{^{ab} e^b_x)\cr}\eqno(15)$$ where a dot denotes time derivative and a prime spatial derivative. The lagrangian (15) has a canonical structure, with coordinates $(e^a_x, \o_x)$, conjugate momenta $(\y_a,\y_2)$ and Lagrange multipliers $(e^a_t, \o_t)$ enforcing the constraints: $$G_a=\y_a'+\e\bam\y_b\o_x+\left({\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x\y_2^h+\gi \y_c\y^c\right)\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r} e_x^b=0$$ $$G_2=\y_2'+\y_a\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r} e^b_x=0\eqno(16)$$ Combining the two constraints (16), one can deduce that $$\ha(\y_a\y^a)'-{\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x\y_2^h\y_2'-\gi\y_a\y^a\y_2'=0\eqno(17)$$ which implies the existence of the conserved quantity $$Q\id\y_a\y^ae^{-2\y_2/\g}-\L\left({\g\over 2}\right)^\hp\G\left(h+1, {2\y_2\over\g}\right)\eqno(18)$$ with $\G$ the incomplete gamma function. The study of the algebra of constraints permits to discuss the symmetries of the theory. The calculation of the Poisson brackets of the constraints yields $$\{G_a,G_2\}=\e\bam G_b\qquad \{G_a,G_b\}=\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r}{\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x h\y_2^{h-1}G_2+\gi\y_cG_c \eqno(19)$$ with coordinate dependent structure functions. This algebra acts locally on the fields by the infinitesimal transformations }\def\ther{thermodynamical }\def\coo{coordinates : $$\d e^a=d\x^a+\e{^a_{\ b}}}\def\bam{{_a^{\ b}}}\def\ba{^{ab}(\x^b\o-\x^2e^b)-\gi\e_{bc}\x^be^c\y^a\qquad\qquad \d\o=d\x^2-{\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x h\y_2^{h-1}}\def\hp{{h+1}}\def\Lq{{\L\over 4}}\def\gi{{1\over\g}\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r}\x^a e^b$$ $$\d\y_a=\e\bam\left[\x^2\y_b+\x_b\left({\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x\y_2^h+\gi\y^a\y_a\right)\right] \qquad\qquad\d\y_2=\e\bam\x^a\y_b$$ as can be checked by computing the commutators $\d e^a=\{G_a,e^b\}$, etc. The lagrangian (13) is invariant under these transformations up to a total derivative. Our model can therefore be considered as a gauge theory of the group generated by the non-linear algebra (19), realized by means of its action on the lagrangian (13). The generalization of the usual gauge theories to non-linear algebras has been introduced in [14], where also the special case $h=2$ of our model has been examined. It must be noticed, however, that in this form the algebra fails to close. In order to construct a closed algebra one has to include in it also the fields $\y_i$ ($i=0,1,2)$ and to consider the family of generators $A(\y_i)+B(\y_i)G_i$, with $A$, $B$ analytic functions of $\y_i$ [11]. One has then: $$\{\y_a,\y_2\}=\{\y_a,\y_b\}=0$$ $$\{G_2,\y_2\}=0\qquad\qquad\{G_a,\y_b\}=\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r}\left({\L\over 2}}\def\gh{{\g\over 2}}\def\edx{\int e\ d^2x\y_2^h+\gi\y_c\y^c \right)\eqno(20)$$ $$\{G_2,\y_a\}=-\{G_a,\y_2\}=\e\bam\y_b$$ The resulting algebra is a nonlinear deformation of $iso(1,2)$ of the kind discussed in [17]. \section{6. Dirac quantization} The model can now be quantized in the Dirac formalism, by replacing the Poisson brackets with commutators and imposing the Gauss law on the physical states. In a momentum representation for the wave functional, $e^a\to i{d\over d\y_a}$, $\o\to i{d\over d\y_2}$, the constraint equations become: $$\left[\y_a'+i\e\bam\y_b{\de\over\de\y_2}+i\e_{ab}}\def\ef{e^{-2\f}}\def\er{e^{-2\r}}\def\erp{e^{2\r}\left(\Lq\y_2^h+{1\over 2\g} \y_c\y^c\right){\de\over\de\y_b}\right]\Q(\y_a,\y_2)=0\eqno(21)$$ $$\left(\y_2'+i\e\bam\y_a{\de\over\de\y_b}\right)\Q(\y_a,\y_2)=0 \eqno(22)$$ The solution of these equations can be written as: $$\Q=\d(Q')e^{i\Omega}\def\P{\Pi}\def\Q{\Psi}\def\S{\Sigma}\def\U{\Upsilon}\def\X{\Xi}\q(Q)\eqno(23)$$ where $Q$ is given in (18) and $$\Omega}\def\P{\Pi}\def\Q{\Psi}\def\S{\Sigma}\def\U{\Upsilon}\def\X{\Xi=\int{\e\ba\y_2\y_ad\y_b\over\y^c\y_c}\eqno(24)$$ The parameters $\L$ and $\g$ enter in (23) only through the parameter $Q$, which classifies the quantum states. Some special cases of the solution (23) have been obtained in [5,13,18]. It should be pointed out, however, that one can not straightforwardly define a Schr\"od- inger equation, since due to the constraints (21,22), the hamiltonian vanishes on the physical states. This is a well-known problem in the hamiltonian quantization of gravity and can be solved by fixing a gauge: in a two-dimensional context it has been treated in ref. [13]. \section{7. Conclusions} We have shown that most of the results obtained in two-dimensional gravity with quadratic curvature and torsion can be extended to the case of an action containing arbitrary powers of the curvature scalar. A possible generalization of these results would be to consider actions containing arbitrary functions of of the curvature, which can be treated essentially by the same methods used here [3,4]. Another interesting point would be to find the most general solutions of the field equations, including time-dependent ones, which we have not considered. It seems plausible that this can be achieved by means of a suitable generalization of the procedure followed in ref. [9] and [10] in the special case of quadratic curvature. \beginref \ref [1] H.J. Schmidt, \JMP{32}, 1562 (1991); \ref [2] S. Mignemi, \PR{D50}, 4733 (1994); \ref [3] S.N. Solodukhin, \PR{D51}, 591 (1995); \ref [4] S. Mignemi and H.-J. Schmidt, \CQG{12}, 845 (1995); \ref [5] S. Mignemi, in press on Ann. Phys.; \ref [6] C. Teitelboim, in {\sl Quantum Theory of gravity}, S.M. Christensen, ed. (Adam Hilger, Bristol, 1984); R. Jackiw, {\sl ibidem}; \ref [7] G. Mandal, A.M. Sengupta and S.R. Wadia, \MPL{A6}, 1685 (1991); \ref [8] D. Cangemi and R. Jackiw, \PRL{69}, 233 (1992); R. Jackiw, Theor. Math. Phys. {\bf 9}, 404 (1992); \ref [9] M.O. Katanaev and I.V. Volovich, \PL{B175}, 413 (1986); \AP{N.Y.} {197}, 1 (1990); M.O. Katanaev, \JMP{31}, 882 (1990); {\bf 32}, 2483 (1991); \ref [10] W. Kummer and D.J. Schwarz, \PR{D45}, 3628 (1992); \ref [11] H.Grosse, W. Kummer, P. Pre\v snajder and D.J. Schwarz, \JMP{33}, 3892 (1992); \ref [12] T. Strobl, \IJMP{A8}, 1383 (1993); \ref [13] T. Strobl, \IJMP{D3}, 281 (1994); \ref [14] N. Ikeda, \AP{N.Y.}{235}, 435 (1994); N. Ikeda and K.I. Izawa, \PTP{89}, 223 (1993); {\bf 89}, 1077 (1993); {\bf 90}, 237 (1993); \ref [15] K.G. Akdeniz, A. Kizilers\"u and A. Rizao\v glu, \PL{B215}, 81 (1988); K.G. Akdeniz, \"O.F. Dayi and A. Kizilers\"u, \MPL{A7}, 1757 (1992); \ref [16] T. Fukuyama and K. Kamimura, \PL{B160}, 259 (1985); K. Isler and C.A. Trugenberger, \PRL{63}, 834 (1989); A.H. Chamseddine and D. Wyler, \PL{B228}, 75 (1989); \ref [17] K. Schoutens, A. Sevrin and P. van Nieuwenhuizen, \CMP{124}, 87 (1989); \ref [18] D. Cangemi and R. Jackiw, \PR{D50}, 3916 (1994). \par\endgroup \end
\section{Introduction} A few years ago, a line of investigation has been proposed by Avdeev and Chizhov \cite{ac} that consists in treating skew-symmetric rank-2 tensor fields as matter rather than gauge degrees of freedom. The model studied in Ref.\cite{ac} has been further reassessed from the point of view of renormalization in the framework of BRS quantization \cite{nos}. In view of the potential relevance of matter-like tensor fields for phenomenology \cite{ac}, it is our purpose in this paper to discuss some facts concerning the formulation of an $N=1$ supersymmetric Abelian gauge model realizing the coupling of gauge fields to matter tensor fields and their partners. One intends here to present a superspace formulation of the model and exploit the possible relevance of extra bosonic supersymmetric partners (complex scalars) for the issue of symmetry breaking. We would like to mention that the supersymmetrization of 2-forms that appear as gauge fields is already known in connection with supergravity, and the so-called linear superfields appear to be the most appropriate multiplets to accomodate the 2-form gauge fields \cite{Gates}. In our case, we aim at the supersymmetrization of matter 2-form fields coupled to Abelian gauge fields and their supersymmetric partners. The present work is outlined as follows: in Section $2$, one searches for the supermultiplet that accomodates the matter tensor field and discusses its self-interaction; the coupling to the gauge supermultiplet is pursued in Sections $3$ and $4$; in Section $5$, one couples the well-known O' Raifeartaigh model \cite{rai} to the tensor-field supermultiplet and discusses some features concerning spontaneous symmetry and supersymmetry breaking. Finally, General Conclusions are drawn in Section $6$. \section{Supersymmetrizing the tensor field} Adopting conventions for the spinor algebra and the superspace parametrization of Ref.\cite{olie}, one finds that the superfield accomodating the skew-symmetric rank-$2$ tensor amongst its components is a spinor multiplet subject to the chirality constraint: \begin{equation}\begin{array}{rl} \S_{a} = & \psi_{a} + \t^{b}\L_{ba} + \t^{2}{\cal F}_{a} - i \t^{c}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{c\dot{c}} \bar{\t}^{\dot{c}} \partial_{\mu} \psi_{a} \\[2mm] & - i \t^{c}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{c\dot{c}} \bar{\t}^{\dot{c}} \partial_{\mu} \t^{b} \L_{ba} - \frac{1}{4} \t^{2} \bar{\t}^{2} \partial_{\mu}\partial^{\mu} \psi_{a} , \end{array}\eqn{sig} \begin{equation}\begin{array}{rl} \overline{\S}_{\dot{a}} = & \overline{\psi}_{\dot{a}} + \bar{\t}_{\dot{b}} \bar{\L}^{\dot{b}}_{\;\,\dot{a}} + \bar{\t}^{2} \overline{{\cal F}}_{\dot{a}} + i \t^{c}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{c\dot{c}} \bar{\t}^{\dot{c}} \partial_{\mu} \overline{\psi}_{\dot{a}} \\[2mm] & + i \t^{c}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{c\dot{c}} \bar{\t}^{\dot{c}} \partial_{\mu} \bar{\t}_{\dot{b}} \bar{\L}^{\dot{b}}_{\;\,\dot{a}} - \frac{1}{4} \t^{2} \bar{\t}^{2} \partial_{\mu}\partial^{\mu} \overline{\psi}_{\dot{a}} , \end{array}\eqn{sigba} \begin{equation}\begin{array}{l} \overline{D}_{\dot{b}}\S_{a} = D_{b}\overline{\S}_{\dot{a}} = 0, \end{array}\eqn{chiralidade} where $\psi_{a}$ and ${\cal F}_{a}$ are chiral spinors and $\L_{ba}$, $\bar{\L}_{\dot{b}\dot{a}}$ are decomposed as: \begin{equation}\begin{array}{l} \L_{ba} = \varepsilon_{ba}\rho + \sigma} \renewcommand{\S}{\Sigma^{{\mu\nu}}_{ba} \l_{{\mu\nu}} \ , \\[2mm] \bar{\L}_{\dot{b}\dot{a}} = - \varepsilon_{\dot{b}\dot{a}} \rho^{\ast} - \overline{\sigma} \renewcommand{\S}{\Sigma}^{{\mu\nu}}_{\dot{b}\dot{a}} \l^{\ast}_{{\mu\nu}} \ . \end{array}\eqn{deflambda} According to the chiral properties of the superfield $\S_{a}$, the $\l_{{\mu\nu}}$-tensor corresponds to the $(1,0)$-representation of Lorentz group. On the other hand, $\l^{\ast}_{{\mu\nu}}$ yields the $(0,1)$-representation. We then write: \begin{equation}\begin{array}{l} \l_{{\mu\nu}} = T_{{\mu\nu}} - i \wti T_{{\mu\nu}} \ , \\[2mm] \l^{\ast}_{{\mu\nu}} = T_{{\mu\nu}} + i \wti T_{{\mu\nu}} \ , \end{array}\eqn{osT} where $\wti T_{{\mu\nu}} = \frac 1 2 \varepsilon_{\mu\nu\a\b} T^{\a\b}$. Notice also that $\wti \l_{{\mu\nu}} = i \l_{{\mu\nu}}$ and $\widetilde{(\l_{{\mu\nu}}^{\ast})} = -i \l_{{\mu\nu}}^{\ast}$, where the twiddle stands for the dual. The canonical dimensions of the component fields read as below: \begin{equation}\begin{array}{l} d(\psi) = d(\overline{\psi}) = \frac{1}{2} \ \\[2mm] d(\rho) = d(\l_{{\mu\nu}}) = 1 \ \\[2mm] d({\cal F}) = d(\overline{{\cal F}}) = \frac{3}{2} \ . \end{array}\eqn{graus} Based on dimensional arguments, we propose the following superspace action for the $\S_{a}$-superfield: \begin{equation} {\cal S} = \int d^4 \! x \, d^{2}\t d^{2}\overline{\t} \hspace{.2cm} \frac{-1}{32}\{ D^{a}\S_{a} \overline{D}_{\dot{a}}\overline{\S}^{\dot{a}} + q\S^{a}\S_{a}\overline{\S}_{\dot{a}}\overline{\S}^{\dot{a}} \}. \eqn{superacao} To check whether such an action is actually the supersymmetric extension of the model that treats $T_{{\mu\nu}}$ as a matter field \cite{ac}, we have now to write down eq.($2.7$) in terms of the component fields $\psi, \rho, \l_{{\mu\nu}}$ and ${\cal F}$: \begin{equation}\begin{array}{rl} {\cal S} & = \int d^4 \! x \, \displaystyle{\Biggl(} + \partial^{\mu}\rho \partial_{\mu}\rho^{\ast} - 16 \partial^{\mu}\l_{{\mu\nu}} \partial_{\a}\l^{\ast\a\nu} + i \overline{{\cal F}}^{\dot{a}} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a} a} \partial_{\mu} {\cal F}^{a} - i \overline{\psi}^{\dot{a}} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a} a} \partial_{\mu} \partial^{\nu}\partial_{\nu} \psi^{a} \\[2mm] & - q \frac{1}{\sqrt{2}}(\rho^{2} - 4 \l_{{\mu\nu}}\l^{{\mu\nu}} ) \frac{1}{\sqrt{2}}(\rho^{\ast 2} - 4 \l^{\ast}_{\a\b}\l^{\ast \a\b} ) +4q \l^{{\mu\nu}}\l_{{\mu\nu}}\overline{{\cal F}}_{\dot{a}}\overline{\psi}^{\dot{a}} +4q \l^{\ast{\mu\nu}}\l_{{\mu\nu}}^{\ast}{\cal F}^{a}\psi_{a} \\[2mm] & - 2q {\cal F}^{a}\psi_{a}\overline{{\cal F}}_{\dot{a}}\overline{\psi}^{\dot{a}} - q\rho^{2}\overline{{\cal F}}_{\dot{a}}\overline{\psi}^{\dot{a}} - q(\rho^{\ast})^{2}{\cal F}^{a}\psi_{a} + \frac{q}{2}\psi^{a}\psi_{a}\partial^{\mu}\partial_{\mu} (\overline{\psi}_{\dot{a}}\overline{\psi}^{\dot{a}}) \\[2mm] & - i q\rho\psi^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\partial_{\mu}(\overline{\psi}^{\dot{a}}\rho^{\ast}) + 4 q \rho\psi^{a}\sigma} \renewcommand{\S}{\Sigma_{a\dot{a}}^{\mu}\partial_{\b}(\l^{\ast\b}_{\:\;\:\;\mu} \overline{\psi}^{\dot{a}}) - 4q \l_{\mu\b}\partial^{\b}(\rho^{\ast} \overline{\psi}_{\dot{a}})\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu \dot{a}a} \psi_{a} \\[2mm] & -16iq \l_{\mu\a}\partial_{\b}(\l^{\ast\b\a}\overline{\psi}_{\dot{a}}) \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu\dot{a} a} \psi_{a} \displaystyle{\Biggr)}. \end{array}\eqn{acaocom} Using $\l_{{\mu\nu}} = \frac{1}{4} ( T_{{\mu\nu}} - i \wti T_{{\mu\nu}} )$, we arrive at the relation: \begin{equation} 16 \partial^{\a}\l_{\a\mu}\partial_{\b}\l^{\ast\b\mu} = 2 \partial^{\a}T_{\a\mu}\partial_{\b}T^{\b\mu} - \frac 1 2 \partial^{\a}T^{{\mu\nu}}\partial_{\a}T_{{\mu\nu}} \; . \eqn{bilinear} The action above displays the terms proposed by Avdeev et al. in Ref.\cite{ac}; besides the anti-symmetric tensor, there appear a complex scalar and a pair of spinors as its supersymmetric partners: $ \psi_{a}$ , a non - physical fermion, and ${\cal F}_{a}$ , corresponding to a physical Weyl spinor. The undesirable presence of a spinor with lower canonical dimension ( $\frac{1}{2}$ , instead of $\frac{3}{2} $ ) generating, as expected, a higher derivative term in the Lagrangian can be avoided by a reshuffling of the spinorial degrees of freedom, if one keeps the interactions turned off. In fact, one can join both $\psi_{a}$ and ${\cal F}_{a}$ in a single fundamental Dirac spinor, $\Psi $, as follows: \begin{equation} \Psi \; = \; \left( \begin{array}{c} {\cal F}_{a}(x) \\ \overline\sigma} \renewcommand{\S}{\Sigma^{\mu\dot{a}a}\partial_{\mu}\; \psi_{a}(x) \end{array} \right) \; . \eqn{OneDirac} The usual kinetic term, $ i\; \overline\Psi\;\gamma^{\mu}\;\partial_{\mu}\; (\Psi) $ , provides the kinetical terms for $\psi_{a}$ and ${\cal F}_{a}$, turning the higher derivative term $, -i\;\overline\psi^{\dot{a}}\overline \sigma} \renewcommand{\S}{\Sigma^{\mu}_{\dot{a}a}\partial_{\mu}\partial^{2} \psi^{a} $, into a matter of choice for the field basis. Nevertheless, this is true only for the free theory. The interaction sector of (\ref{acaocom}) cannot be re-expressed in terms of the Dirac spinor $\Psi $ , imposing a dissociation back to Weyl spinorial degrees of freedom. Therefore, it happens that the full theory must carry a higher derivative term, giving birth to a conjecture that this might be the fermionic counterpart of problems concerning the transverse sector of the original - bosonic - model for the tensor matter field with interactions, as discussed in \cite{ac}. \section{The gauging of the model} In order to perform the gauging of the model described by eq.(2.7), one proceeds along the usual lines and introduces a chiral scalar superfield, $\L$, to act as the gauge parameter: \begin{equation} \L = ( 1 - i \t^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\t}^{\dot{a}}\partial_{\mu} - \frac{1}{4} \t^{2} \overline{\t}^{2} \partial_{\mu}\partial^{\mu} ) ( \phi + \t^{b}w_{b} + \t^{2}\pi ) \eqn{invsemb} \begin{equation} \overline{\L} = ( 1 + i \t^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\t}^{\dot{a}}\partial_{\mu} - \frac{1}{4} \t^{2} \overline{\t}^{2} \partial_{\mu}\partial^{\mu} ) ( \phi^{\ast} + \overline{\t}_{\dot{b}}\overline{w}^{\dot{b}} + \overline{\t}^{2} \overline{\pi} ). \eqn{invcomb} The infinitesimal gauge transformations of the superfields $\S$ and $\overline{\S}$ read as: \begin{equation}\begin{array}{rl} \d \S_{a} &= ih \L \hspace{.1cm}\S_{a} \\[2mm] \d \overline\S_{\dot{a}} &= -ih \overline{\L}\hspace{.1cm} \overline{\S}_{\dot{a}}, \end{array}\eqn{invsigma} and the behaviour of $(D^{a}\S_{a})$ and $(\overline{D_{\dot{a}}}\hspace{.1cm}\overline{\S^{\dot{a}}})$ under finite transformations become: \begin{equation}\begin{array}{rl} D^{a}\hspace{.1cm}\S_{a}^{'} &= e^{ih\L}\hspace{.1cm}(D^{a}\hspace{.1cm}\S_{a} + ih D^{a}\L\hspace{.1cm} \S_{a}) \\[2mm] \overline{D_{\dot{a}}}\hspace{.1cm}\overline{\S^{'\dot{a}}} &= e^{-ih\overline{\L}}(\overline{D_{\dot{a}}}\hspace{.1cm}\overline{\S^{\dot{a}}} -ih\overline{D_{\dot{a}}}\hspace{.1cm}\overline{\L}\hspace{.1cm} \overline{\S^{\dot{a}}}). \end{array}\eqn{noninv} To gauge-covariantize the superspace derivatives, one introduces a gauge connection superfield: \begin{equation} D_{a} \rightarrow \nabla_{a} = D_{a} + ih \Gamma_{a}, \eqn{dericova} in such a way that $\Gamma_{a}$ transforms like \begin{equation} \Gamma_{a}^{'} = \Gamma_{a} - D_{a}\L . \eqn{invgamma} This yields: \begin{equation} (\nabla^{a}\S_{a})^{'} = e^{ih\L} (\nabla^{a}\S_{a}). \eqn{superinv} To achieve a U$(1)$-invariant action, one proposes \begin{equation} {\cal S} = \int d^4 \! x \, d^{2}\t d^{2}\overline{\t} \displaystyle{\Biggl(} \nabla^{a}\S_{a}\hspace{.1cm}e^{hV}\hspace{.1cm}\overline{\nabla_{\dot{a}}} \overline{\S^{\dot{a}}} \displaystyle{\Biggr)}, \eqn{gaugeaction} where V \hspace{1mm} is the real scalar superfield \cite{jose} that accomplishes the gauging of supersymmetric QED \cite{jose2}: \begin{equation} V^{'} = V + i\hspace{.1cm}( \overline{\L} - \L ). \eqn{invV} At this point, the gauge sector displays more degrees of freedom than what is actually required to perform the gauging. There are component vector fields in both $\Gamma_{a}$ and $V$. However, we notice that the superfield $\Gamma_{a}$ is not a true independent gauge potential. Indeed, \begin{equation} \Gamma_{a} = - i D_{a} V \eqn{relacomv} reproduces correctly the gauge tranformation of $\Gamma_{a}$ and, at the same time, eliminates the redundant degrees of freedom that would be otherwise present, if we were to keep $\Gamma_{a}$ and $V$ as gauge superfields. Therefore, the locally U$(1)$- invariant action takes over the form: \begin{equation}\begin{array}{rl} {\cal S} & = \int d^4 \! x \, d^{2}\t \hspace{.2cm} \frac{-1}{128} \displaystyle{\Biggl(} \overline{D}^{2}(e^{-V} D^{a} e^{V})\overline{D}^{2}(e^{-V} D_{a} e^{V}) \displaystyle{\Biggr)} \\[2mm] & + \int d^4 \! x \, d^{2}\t d^{2}\overline{\t} \hspace{.2cm}\frac{-1}{32} \displaystyle{\Biggl(} \nabla^{a}\S_{a}\hspace{.1cm}e^{hV}\hspace{.1cm} \overline{\nabla_{\dot{a}}}\overline{\S^{\dot{a}}} + q \S^{a}\S_{a}e^{2hV}\overline{\S}_{\dot{a}}\overline{\S}^{\dot{a}} \displaystyle{\Biggr)}, \end{array}\eqn{gaugeaction} where \begin{equation}\begin{array}{rl} \nabla^{a}\S_{a} &= D^{a}\S_{a} + hD^{a}V \hspace{.1cm}\S_{a} \\[2mm] \hspace{2mm} \overline{\nabla_{\dot{a}}}\overline{\S^{\dot{a}}} &= \overline{D_{\dot{a}}}\overline{\S^{\dot{a}}} + h\overline{D_{\dot{a}}} V\hspace{.1cm}\overline{\S^{\dot{a}}}. \end{array}\eqn{superderi} The $\t$-expansion for the superfield $V$ brings about the following component fields: \begin{equation}\begin{array}{rl} V &= C + \t^{a}b_{a} + \overline{\t}_{\dot{a}}\overline{b}^{\dot{a}} + \t^{a} \overline{\t}^{\dot{a}}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}A_{\mu} \\[2mm] & + \hspace{.1cm}\t^{2}\l + \overline{\t}^{2}\overline{\l} + \t^{2}\overline{\t}_{\dot{a}}\overline{\gamma}^{\dot{a}} + \overline{\t}^{2}\t^{a}\gamma_{a} + \t^{2}\overline{\t}^{2}\Delta, \end{array}\eqn{expancao} where C, $\l$, $\overline{\l}$ and $\Delta$ are scalars $b_{a}$ and $\gamma_{a}$ are spinors and $A_{\mu}$ is the U($1$)-gauge field. The gauge transformation of these fields read as below: \begin{equation}\begin{array}{rl} \d C &= i( \phi^{\ast} - \phi) ,\hspace{.2cm} \d\l = -i\pi,\hspace{.2cm} \d\overline{\l} = i\overline{\pi},\hspace{.2cm} \\[2mm] \d b_{a} &= -iw_{a},\hspace{.2cm} \d \overline{b}_{\dot{a}} = i\overline{w}_{\dot{a}} \\[2mm] \d\Delta &= \frac{i}{4}\partial^{\mu}\partial_{\mu} (\phi - \phi^{\ast}),\hspace{.2cm} \d A^{\mu} = -\partial^{\mu}(\phi + \phi^{\ast}),\hspace{.2cm} \\[2mm] \d\gamma_{a} &= \frac{1}{2}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\partial_{\mu}\overline{w}^{\dot{a}},\hspace{.2cm} \d\overline{\gamma}_{\dot{a}} = - \frac{1}{2}\overline{\sigma} \renewcommand{\S}{\Sigma}_{\dot{a}a}\partial_{\mu}w^{a}. \end{array}\eqn{compinv} As already known, for the sake of component-field calculations, one usually works in the so-called Wess-Zumino gauge, where C, $b_{a}$ and $\l$ are gauged away. The expansion for the exponential of the gauge superfield simplifies, in this gauge, according to: \begin{equation} e^{hV} = 1 + h\t^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\t}^{\dot{a}} A_{\mu} + h\t^{2}\overline{\t}_{\dot{a}}\overline{\gamma}^{\dot{a}} + h\overline{\t}^{2}\t^{a}\gamma_{a} + h\t^{2}\overline{\t}^{2}\Delta + \frac{1}{4} h^{2}\t^{2}\overline{\t}^{2}A^{\mu}A_{\mu}. \eqn{expe} Using this gauge, the transformations of the matter fields are: \begin{equation}\begin{array}{rl} \psi_{a} &= ih\phi\psi_{a},\hspace{.2cm} \d\rho = ih\phi\rho,\hspace{.2cm} \d\l_{{\mu\nu}} = ih\phi\l_{{\mu\nu}},\hspace{.2cm} \d{\cal F}_{a} = ih( \phi{\cal F}_{a} ); \end{array}\eqn{invmatter} we get thereby the following transformations for the components $T_{{\mu\nu}}$ and $\wti T_{{\mu\nu}}$: \begin{equation} \d T_{{\mu\nu}} = h\phi \wti T_{{\mu\nu}}, \hspace{.5cm}\d \wti T_{{\mu\nu}} = - h\phi T_{{\mu\nu}} \eqn{invqui} These are precisely the Abelian gauge transformations for the tensor field as firstly proposed in Ref.\cite{ac}. \section{Component-field action in the Wess-Zumino gauge} Having adopted the component fields as defined in the previous sections, lengthy algebraic computations yield the following action in the Wess-Zumino gauge: \begin{displaymath}\begin{array}{rl} {\cal S} &= \int d^4 \! x \, \displaystyle{\Biggl(} -\frac{1}{4}F^{{\mu\nu}}F_{{\mu\nu}} +2 \Delta^{2} + i \overline{\gamma}^{\dot{a}}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a}a} \partial_{\mu}\gamma^{a} + \partial^{\mu}\rho\partial_{\mu}\rho^{\ast} - 16 \partial^{\mu}\l_{{\mu\nu}} \partial_{\a}\l^{\ast\a\nu} \\[2mm] & + i \overline{{\cal F}}^{\dot{a}} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a} a} \partial_{\mu} {\cal F}^{a} - i \overline{\psi}^{\dot{a}} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a} a} \partial_{\mu} \partial^{\nu}\partial_{\nu} \psi^{a} - i\frac{h}{2}\partial^{\mu}\rho A_{\mu}\rho^{\ast} + i\frac{h}{2} \rho A^{\mu}\partial_{\mu}\rho^{\ast} \\[2mm] & + 2h\rho\Delta\rho^{\ast} + \frac{h^{2}}{4}\rho A^{\mu}A_{\mu}\rho^{\ast} - h\gamma^{a}{\cal F}_{a}\rho^{\ast} - h\rho\overline{\gamma}_{\dot{a}}\overline{{\cal F}}^{\dot{a}} - i\frac{h}{2}\rho\gamma^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\partial_{\mu}\overline{\psi}^{\dot{a}} \\[2mm] & -i\frac{h}{2} \rho^{\ast}\overline{\gamma}^{\dot{a}}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a}a}\partial_{\mu}\psi^{a} + i \frac{h}{2} \partial_{\mu}\rho^{\ast}\psi^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\gamma}^{\dot{a}} - i\frac{h}{2} \partial_{\mu}\rho\gamma^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\psi}^{\dot{a}} - \frac{h}{4} \psi^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}A_{\mu}\partial^{\nu}\partial_{\nu}\overline{\psi}^{\dot{a}} \\[2mm] & + \frac{h}{4} \overline{\psi}^{\dot{a}}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a}a}A_{\mu}\partial^{\nu} \partial_{\nu}\psi^{a} + i h \Delta\overline{\psi}^{\dot{a}}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a}a} \partial_{\mu}\psi^{a} + i h \Delta\psi^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\partial_{\mu}\overline{\psi}^{\dot{a}} -\frac{h}{4}\partial^{\nu}A_{\nu}\overline{\psi}^{\dot{a}}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a}a} \partial_{\mu}\psi^{a} \\[2mm] & + \frac{h}{4}\partial_{\nu}A^{\nu}\psi^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\partial_{\mu} \overline{\psi}^{\dot{a}} + \frac{h}{2} {\cal F}^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}A_{\mu} \overline{{\cal F}}^{\dot{a}} - \frac{h}{8}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\nu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu} \sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}} F_{\mu\nu}\partial_{\a}\overline{\psi}^{\dot{a}} \\[2mm] &+ \frac{h}{8}\partial_{\mu}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\mu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\a}\sigma} \renewcommand{\S}{\Sigma^{\nu})_{a\dot{a}} F_{\nu\a}\overline{\psi}^{\dot{a}} + h \rho^{\ast}F_{{\mu\nu}}\l^{{\mu\nu}} + h\rho F_{\mu\nu}\l^{\ast\mu\nu} \\[2mm] & + 2ih ( 4 \partial_{\mu}\l^{\ast\mu\nu}\l_{\nu\a}A^{\a} - 4 \partial^{\mu}\l_{{\mu\nu}}A_{\a}\l^{\ast\nu\a}) \\[2mm] &- i \frac{h^{2}}{2}\gamma^{a}(\sigma} \renewcommand{\S}{\Sigma^{\mu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\nu}\sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}} \l_{{\mu\nu}}A_{\a}\overline{\psi}^{\dot{a}} + \frac{h}{2}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\nu} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}\sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}} \overline{\gamma}^{\dot{a}}\partial_{\nu} \l^{\ast}_{\mu\a} - \frac{h}{2}\partial_{\nu}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\nu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}\sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}} \l^{\ast}_{\mu\a} \overline{\gamma}^{\dot{a}}\\[2mm] &+ i\frac{h^{2}}{2} \overline{\gamma}^{\dot{a}}(\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}\sigma} \renewcommand{\S}{\Sigma^{\a} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\b})_{\dot{a}a}A_{\mu}\psi^{a}\l^{\ast}_{\a\b} - \frac{h}{2} \gamma^{a}(\sigma} \renewcommand{\S}{\Sigma^{\a}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\b}\sigma} \renewcommand{\S}{\Sigma^{\mu})_{a\dot{a}}\partial_{\mu}\l_{\a\b} \overline{\psi}^{\dot{a}} + \frac{h}{2}\gamma^{a}(\sigma} \renewcommand{\S}{\Sigma^{a} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\b}\sigma} \renewcommand{\S}{\Sigma^{\mu})_{a\dot{a}}\l_{\a\b}\partial_{\mu}\overline{\psi}^{\dot{a}} \\[2mm] & + \frac{h}{2}\partial_{\mu}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\mu} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\nu}\sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}}A_{\nu}\partial_{\a}\overline{\psi}^{\dot{a}} + \frac{h}{4}\partial_{\mu}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\nu} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}\sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}}A_{\nu}\partial_{\a}\overline{\psi}^{\dot{a}} \\[2mm] &+ \frac{h}{4} \partial_{\mu}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\mu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\a}\sigma} \renewcommand{\S}{\Sigma^{\nu})_{a\dot{a}}A_{\nu}\partial_{\a} \overline{\psi}^{\dot{a}} + h^{2}\Delta\psi^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}A_{\mu} \overline{\psi}^{\dot{a}} +\frac{i}{16}h^{2}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\nu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu} \sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}} F_{{\mu\nu}}A_{\a}\overline{\psi}^{\dot{a}} \\[2mm] & + i \frac{h^{2}}{16}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\nu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}\sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}} F_{\a\mu}A_{\nu}\overline{\psi}^{\dot{a}} -i\frac{h^{2}}{2}\gamma^{a}(\sigma} \renewcommand{\S}{\Sigma^{\a} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\b}\sigma} \renewcommand{\S}{\Sigma^{\mu})_{a\dot{a}} \l_{\a\b}A_{\mu}\overline{\psi}^{\dot{a}} \; + \end{array}\end{displaymath} \begin{equation}\begin{array}{rl} & - i\frac{h^{2}}{4}\partial_{\mu}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\mu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\nu} \sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}}A_{\nu}A_{\a}\overline{\psi}^{\dot{a}} - i \frac{h^{2}}{8}\partial_{\mu}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\nu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu} \sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}}A_{\nu}A_{\a}\overline{\psi}^{\dot{a}} \\[2mm] & + i \frac{h^{2}}{4} \psi^{a}A_{\nu}A_{\mu}(\sigma} \renewcommand{\S}{\Sigma^{\nu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}\sigma} \renewcommand{\S}{\Sigma^{\a})_{a\dot{a}} \partial_{\a}\overline{\psi}^{\dot{a}} + i \frac{h^{2}}{8}\psi^{a}A_{\nu}A_{\mu} (\sigma} \renewcommand{\S}{\Sigma^{\nu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\a} \sigma} \renewcommand{\S}{\Sigma^{\mu})_{a\dot{a}}\partial_{\a}\overline{\psi}^{\dot{a}} \\[2mm] & + \frac{h^{3}}{8}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\nu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\a}\sigma} \renewcommand{\S}{\Sigma^{\mu})_{a\dot{a}} A_{\nu}A_{\a}A_{\mu}\overline{\psi}^{\dot{a}} - h^{2}\gamma^{a}\psi_{a} \overline{\gamma}_{\dot{a}}\overline{\psi}^{\dot{a}} + 4 h^{2}A^{\nu}A_{\mu} \l_{\nu\b}\l^{\ast\b\mu} \\[2mm] &- q \frac{1}{\sqrt{2}}(\rho^{2} - 4 \l_{{\mu\nu}}\l^{{\mu\nu}} ) \frac{1}{\sqrt{2}}(\rho^{\ast 2} - 4 \l^{\ast}_{\a\b}\l^{\ast \a\b} ) +4q \l^{{\mu\nu}}\l_{{\mu\nu}}\overline{{\cal F}}_{\dot{a}}\overline{\psi}^{\dot{a}} +4q \l^{\ast{\mu\nu}}\l_{{\mu\nu}}^{\ast}{\cal F}^{a}\psi_{a} \\[2mm] & - 2q {\cal F}^{a}\psi_{a}\overline{{\cal F}}_{\dot{a}}\overline{\psi}^{\dot{a}} - q\rho^{2}\overline{{\cal F}}_{\dot{a}}\overline{\psi}^{\dot{a}} - q(\rho^{\ast})^{2}{\cal F}^{a}\psi_{a} + \frac{q}{2}\psi^{a}\psi_{a}\partial^{\mu}\partial_{\mu} (\overline{\psi}_{\dot{a}}\overline{\psi}^{\dot{a}}) \\[2mm] &- iq\rho\psi^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\partial_{\mu}(\overline{\psi}^{\dot{a}}\rho^{\ast}) + 4q \rho\psi^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\partial^{\b}(\l^{\ast}_{\b\mu}\overline{\psi}^{\dot{a}}) - 4q \l_{\mu\b}\partial^{\b}(\rho^{\ast}\overline{\psi}_{\dot{a}}) \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu\dot{a}a}\psi_{a} \\[2mm] &- 16i q \l_{\mu\a} \partial_{\b}(\l^{\ast\b\a}\overline{\psi}_{\dot{a}}) \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu\dot{a}a}\psi_{a} - q h \Delta\psi^{a}\psi_{a} \overline{\psi}_{\dot{a}}\overline{\psi}^{\dot{a}} - iq\frac{h}{2} A_{\mu}\psi^{a}\psi_{a}\partial^{\mu}(\overline{\psi}_{\dot{a}} \overline{\psi}^{\dot{a}}) \\[2mm] & + iq\frac{h}{2} A_{\mu}\partial^{\mu}(\psi^{a}\psi_{a})\overline{\psi}_{\dot{a}}\overline{\psi}^{\dot{a}} - q \frac{h^{2}}{2} A^{\mu}A_{\mu}\psi^{a}\psi_{a}\overline{\psi}_{\dot{a}}\overline{\psi}^{\dot{a}} - q h \rho\gamma^{a}\psi_{a}\overline{\psi}_{\dot{a}}\overline{\psi}^{\dot{a}} - q h \rho^{\ast}\overline{\gamma}_{\dot{a}}\overline{\psi}^{\dot{a}}\psi^{a}\psi_{a} \\[2mm] &- q h\gamma^{a}\sigma} \renewcommand{\S}{\Sigma^{{\mu\nu}}_{ab}\psi^{b}\overline{\psi}_{\dot{a}} \overline{\psi}^{\dot{a}}\l_{{\mu\nu}} + q h \overline{\gamma^{\dot{a}}} \overline{\sigma} \renewcommand{\S}{\Sigma}^{{\mu\nu}}_{\dot{a}\dot{b}}\l^{\ast}_{{\mu\nu}} \overline{\psi}^{\dot{b}}\psi^{a}\psi_{a} - q h \rho\psi^{a} \sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}A_{\mu}\overline{\psi}^{\dot{a}}\rho^{\ast} \\[2mm] & - 2i q h \rho\psi^{a}A_{\mu}(\sigma} \renewcommand{\S}{\Sigma^{\mu}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\a} \sigma} \renewcommand{\S}{\Sigma^{\b})_{a\dot{a}}\l^{\ast}_{\a\b}\overline{\psi}^{\dot{a}} + 2 i q h \rho^{\ast}\psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\b}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\a}\sigma} \renewcommand{\S}{\Sigma^{\mu})_{a\dot{a}} A_{\mu}\l_{\a\b}\overline{\psi}^{\dot{a}} \\[2mm] & + 2 q h \psi^{a}(\sigma} \renewcommand{\S}{\Sigma^{\a} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\b}\sigma} \renewcommand{\S}{\Sigma^{\gamma} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}\sigma} \renewcommand{\S}{\Sigma^{\nu})_{a\dot{a}}\l_{\a\b}A_{\gamma}\l^{\ast}_{\mu\nu} \overline{\psi}^{\dot{a}} \displaystyle{\Biggr)}. \end{array}\eqn{wsacao} We should stress here a remarkable difference with respect to the case of the chiral and anti-chiral \underline{scalar} superfields (Wess-Zumino model \cite{WZM}), namely, the minimal coupling of $\S$ and $\overline{\S}$ to the gauge sector necessarily affects the $\S$ - superfield self-interaction terms as one reads off from eq ($3.11$). The gauging of the $U(1)$ - symmetry enriches the self-interactions of the tensor field not only through its fermionic supersymmetric partners, but also through the introduction of the gauge boson and the gaugino at the level of the matter self-interaction terms (these new couplings are compatible with power-counting renormalizability). This is so because the model presented here is based on a single spinor superfield. Had we introduced a couple of spinor superfields, $\S_{a}$ and ${\cal T}_{a}$, with opposite $U(1)$ charges: \begin{equation}\begin{array}{rl} \S^{'}_{a} &= e^{ih\L}\S_{a}, \\[2mm] {\cal T}^{'}_{a} &= e^{-ih\L}{\cal T}_{a}, \end{array}\eqn{taus} a self-interacting term of the form $(\S^{a}{\cal T}_{a}\overline{\S}_{\dot{a}}\overline{{\cal T}}^{\dot{a}})$ would automatically be invariant whenever the symmetry is gauged, and there would be no need for introducing the vector superfield to ensure local invariance. Such a mixed self-interacting term could, in principle, be thought of as a possible source for a mass term for the spinor superfields, whenever the physical scalar component $\rho$ develops a non-trivial vacuum expectation value. Nevertheless, by analysing the $\rho$ - field interactions in the scalar potential, one concludes that there is no room for spontaneous symmetry breaking as induced by such a component field (and, similarly, for its counterpart inside ${\cal T}$ ). On the other hand, we could think to introduce a gauge-invariant mass term of the form \begin{equation} {\cal S}_{mass} = \int d^4 \! x \, (d^{2}\t \hspace{.2cm}i\frac{m}{16}\S^{a}{\cal T}_{a} - d^{2}\overline{\t} \hspace{.2cm}i\frac{m}{16} \overline{\S}_{\dot{a}}\overline{{\cal T}}^{\dot{a}}); \eqn{qusemass} however, a mixed mass term like the one above introduces two massive excitations of the type $k^{4} = m^{4}$ in the spectrum. So, regardless the sign of $m^{2}$, a tachyon shall always be present; hence such a mass term is disregarded. \section{On Spontaneous Symmetry Breaking} Due to the spinorial character of the superfield $\S_{a}$, it cannot be used to accomplish a spontaneous breaking. Indeed, Lorentz invariance is lost whenever $\S_{a}$ acquires a non-trivial vacuum expectation value. The idea in the present section is to couple, in a gauge-invariant manner, the well-known O' Raifeartaigh model \cite{rai} to the spinor superfield $\S_{a}$, so as to understand the issue of an eventual mass generation for $\S_{a}$ via spontaneous internal symmetry or supersymmetry breakingdown. For the sake of concreteness, the model we adopt to discuss such a matter is defined by the action below: \begin{equation}\begin{array}{rl} {\cal S} &= \int d^4 \! x \, d^{2}\t d^{2}\overline{\t} \displaystyle{\Biggl(} \overline{\phi}\phi + \overline{\phi_{+}}e^{hV}\phi_{+} + \overline{\phi_{-}}e^{-hV}\phi_{-} \displaystyle{\Biggr)} \\[2mm] &+ \int d^4 \! x \, d^{2}\t \displaystyle{\Biggl(} \frac 1 2 m\phi^{2} + \mu \phi_{+}\phi_{-} + f \phi + g \phi\phi_{+}\phi_{-} + G\S^{a}\S_{a}\phi_{-}\fme \displaystyle{\Biggr)} \\[2mm] &+ \int d^4 \! x \, d^{2}\overline{\t} \displaystyle{\Biggl(} \frac 1 2 m\overline{\phi}^{2} + \mu \overline{\phi_{+}} \overline{\phi_{-}} + f \overline{\phi} + g \overline{\phi} \overline{\phi_{+}} \overline{\phi_{-}} + G\overline{\S}_{\dot{a}} \overline{\S}^{\dot{a}}\overline{\phi_{-}} \overline{\phi_{-}} \displaystyle{\Biggr)} \\[2mm] &+ \int d^4 \! x \, d^{2}\t \hspace{.2cm} \frac{-1}{128} \displaystyle{\Biggl(} \overline{D}^{2}(e^{-V} D^{a} e^{V})\overline{D}^{2}(e^{-V} D_{a} e^{V}) \displaystyle{\Biggr)} \\[2mm] & + \int d^4 \! x \, d^{2}\t d^{2}\overline{\t} \hspace{.2cm}\frac{-1}{32} \displaystyle{\Biggl(} \nabla^{a}\S_{a}\hspace{.1cm}e^{hV}\hspace{.1cm}\overline{\nabla_{\dot{a}}} \overline{\S^{\dot{a}}} + q \S^{a}\S_{a}e^{2hV}\overline{\S}_{\dot{a}} \overline{\S}^{\dot{a}} \displaystyle{\Biggr)}, \end{array}\eqn{oraios} where the chiral scalar superfields $\phi$, $\phi_{+}$ and $\phi_{-}$ are parametrized as follows: \begin{equation}\begin{array}{rl} \phi &= ( 1 - i \t^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\t}^{\dot{a}}\partial_{\mu} - \frac{1}{4} \t^{2} \overline{\t}^{2} \partial_{\mu}\partial^{\mu} ) ( A + \t^{b}\xi_{b} + \t^{2}b ) \\[2mm] \phi_{+} &= ( 1 - i \t^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\t}^{\dot{a}}\partial_{\mu} - \frac{1}{4} \t^{2} \overline{\t}^{2} \partial_{\mu}\partial^{\mu} ) ( A_{+} + \t^{b}\xi_{+b} + \t^{2}b_{+} ) \\[2mm] \phi_{-} &= ( 1 - i \t^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\t}^{\dot{a}}\partial_{\mu} - \frac{1}{4}\t^{2} \overline{\t}^{2} \partial_{\mu}\partial^{\mu} ) ( A_{-} + \t^{b}\xi_{-b} + \t^{2}b_{-} ) \; ; \end{array}\eqn{fis} m and $\mu$ are mass parameters, f has dimension of ${mass}^{2}$, whereas g and G are dimensionless coupling constants. $\S_{a}$ and $\phi_{-}$ have opposite U$(1)$ - charges. This action, in terms of components, reads: \begin{displaymath}\begin{array}{rl} {\cal S} &= \int d^4 \! x \, 4\displaystyle{\Biggl(} 4\{ \partial^{\mu}A^{\ast}\partial_{\mu}A + \partial^{\mu}A^{\ast}_{-}\partial_{\mu}A_{-} + \partial^{\mu}A^{\ast}_{+}\partial_{\mu}A_{+} \} + 4\{ b^{\ast}b + b^{\ast}_{-}b_{-} + b^{\ast}_{+}b_{+} \} \\[2mm] &+ 2i \{ \overline{\xi}^{\dot{a}} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a}a} \partial_{\mu} \xi^{a} + \overline{\xi}^{\dot{a}}_{-} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a}a} \partial_{\mu} \xi^{a}_{-} + \overline{\xi}^{\dot{a}}_{+} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{a}a} \partial_{\mu} \xi^{a}_{+} \} \displaystyle{\Biggr)} \; + \end{array}\end{displaymath} \begin{equation}\begin{array}{rl} &+ \int d^4 \! x \, \displaystyle{\Biggl(} 16h A^{\ast}_{+}\Delta A_{+} + 8ih A^{\ast}_{+}\partial^{\mu}A_{\mu}A_{+} + 4h^{2} A^{\ast}_{+}A^{\mu}A_{\mu}A_{+} - 8h A^{\ast}_{+}\gamma^{a}\xi_{+a} \\[2mm] &- 8h\overline{\xi}_{+\dot{a}}\overline{\gamma}^{\dot{a}}A_{+} + 4h \xi^{+a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}A_{\mu}\overline{\xi}^{\dot{a}}_{+} + 16ih\partial^{\mu}A^{\ast}_{+}A_{\mu}A_{+} \displaystyle{\Biggr)} \\[2mm] &+ \int d^4 \! x \, \displaystyle{\Biggl(} -16h A^{\ast}_{-}\Delta A_{-} - 8ih A^{\ast}_{-}\partial^{\mu}A_{\mu}A_{-} + 4h^{2}A^{\ast}_{-}A^{\mu}A_{\mu}A_{-} + 8h A^{\ast}_{-}\gamma^{a}\xi_{-a} \\[2mm] &+ 8h\overline{\xi}_{-\dot{a}}\overline{\gamma}^{\dot{a}}A_{-} - 4h \xi^{-a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}A_{\mu}\overline{\xi}^{\dot{a}}_{-} - 16ih\partial^{\mu}A^{\ast}_{-}A_{\mu}A_{-} \displaystyle{\Biggr)} \\[2mm] &+ \int d^4 \! x \, \displaystyle{\Biggl(} m(-4bA + \xi^{a}\xi_{a}) + m(-4b^{\ast}A^{\ast} + \overline{\xi}_{\dot{a}}\overline{\xi}^{\dot{a}}) - 4fb - 4fb^{\ast} \displaystyle{\Biggr)} \\[2mm] & + \int d^4 \! x \, 2\displaystyle{\Biggl(} \mu(-2b_{+}A_{-} - 2b_{-}A_{+} + \xi^{a}_{+}\xi_{a -}) + \mu(-2b^{\ast}_{+}A^{\ast}_{-} - 2b^{\ast}_{-}A^{\ast}_{+} + \overline{\xi}_{\dot{a}+}\overline{\xi}^{\dot{a}}_{-}) \displaystyle{\Biggr)} \\[2mm] & + \int d^4 \! x \, 2\displaystyle{\Biggl(} g(-2bA_{+}A_{-} - 2Ab_{+}A_{-} - 2AA_{+}b_{-} + A\xi^{a}_{+}\xi_{a -} + A_{-}\xi^{a}\xi_{a +} + A_{+}\xi^{a}\xi_{a -}) \\[2mm] & + g (-2b^{\ast}A^{\ast}_{+}A^{\ast}_{-} - 2A^{\ast}b^{\ast}_{+}A^{\ast}_{-} - 2A^{\ast}A^{\ast}_{+}b^{\ast}_{-} + A^{\ast}\overline{\xi}_{\dot{a} +}\overline{\xi}^{\dot{a} -} + A^{\ast}_{-}\overline{\xi}_{\dot{a}}\overline{\xi}^{\dot{a}}_{+} + A^{\ast}_{+}\overline{\xi}_{\dot{a}}\overline{\xi}^{\dot{a}}_{-}) \displaystyle{\Biggr)} \\[2mm] & + \int d^4 \! x \, 4G\displaystyle{\Biggl(} (-2{\cal F}^{a}\psi_{a} + \rho^{2} +4 \l_{{\mu\nu}}\l^{{\mu\nu}})(A_{-})^{2} + (-2\overline{{\cal F}}_{\dot{a}}\overline{\psi}^{\dot{a}} + (\rho^{\ast})^{2} +4 \l^{\ast}_{{\mu\nu}}\l^{\ast {\mu\nu}})(A^{\ast}_{-})^{2} \\[2mm] & + 2\psi^{b}\psi_{b}(-4b_{-}A_{-} + \xi^{a}_{-}\xi_{a -}) + 2\overline{\psi}_{\dot{b}}\overline{\psi}^{\dot{b}}(-4b^{\ast}_{-}A^{\ast}_{-} + \overline{\xi}_{\dot{a} -}\overline{\xi}^{\dot{a} -}) \\[2mm] & - 2(\rho\psi^{b} - \sigma} \renewcommand{\S}{\Sigma^{{\mu\nu}\,ba}\l_{{\mu\nu}}\psi_{a})(\xi_{b -}A_{-}) +2(\rho^{\ast}\overline{\psi}_{\dot{b}} - \l^{\ast}_{{\mu\nu}}\overline{\sigma} \renewcommand{\S}{\Sigma}^{{\mu\nu}}_{\dot{b}\dot{a}} \overline{\psi}^{\dot{a}})(\overline{\xi}^{\dot{b}}_{-} A^{\ast}_{-}) \displaystyle{\Biggr)} \\[2mm] & + \int d^4 \! x \, \displaystyle{\Biggl(} + 2h\rho\Delta\rho^{\ast} +2 \Delta^{2} + \frac{h^{2}}{4}\rho A^{\mu}A_{\mu}\rho^{\ast} + \ldots \displaystyle{\Biggr)} , \end{array} \eqn{bAxi2} where the dots stand for spinorial partners and derivative terms that are completely irrelevant for discussing spontaneous symmetry breaking. Also, the $\rho $-field does not acquire a non-trivial v.e.v., as already mentioned in the previous section. The scalars that could, in principle, trigger spontaneous symmetry breaking are only $ A $, $ A_{+} $ and $ A_{-} $, if one starts from action (\ref{bAxi2}). The only possible way to endow the tensor field $ \l_{\mu\nu} $ with a mass, via internal symmetry or supersymmetry breaking, would be by means of a coupling of the form \begin{equation} \S^{a}\S_{a} \phi^{2} \; , \eqn{acopla} as dictated by supersymmetry, through the chirality constraints on $ \S $ and $ \phi $. No matter the number of scalar superfields present in the model, whenever the breaking takes place and a mass is generated for $ \l_{\mu\nu} $ as a byproduct, one notices that both the imaginary part of $ \rho $ and one longitudinal degree of freedom of $ \l_{\mu\nu} $ provide the spectrum with a tachyonic excitation, without any chance of avoiding this fact at the expenses of the $\Delta $-field coupling, enriched by an additional Fayet-Iliopoulos term \cite{Fayet} (actually, a $\Delta $-type term does not break supersymmetry whenever it is addded to the O' Raifeartaigh model. Moreover, a $\Delta $-term never couples to $ \l_{\mu\nu} $, since $ \l_{\mu\nu} \l^{\ast}_{\mu\nu} $ is identically vanishing). Therefore, our final conclusion is that the masslessness of $ \rho $ and $ \l_{\mu\nu} $ cannot be avoided in a consistent way, just by invoking internal symmetry or supersymmetry breaking as realised by a set of scalar superfields. \section{General Conclusions} The supersymmetrization of the matter tensor field first investigated in Ref.{\cite{ac}} has been worked out here in terms of a spinor chiral superfield, $\S_{a}$, whose kinetic and self-interacting terms have been found in $N=1$ - superspace. The gauging of the model reveals some peculiarities, such as the need of gauge fields to appear in the matter self-interactions. Scalar degrees of freedom that accompany the fermionic partners of $ \l_{\mu\nu} $ cannot be the source for spontaneous symmetry or supersymmetry breaking, as it could in principle be thought. The reason is that Lorentz invariance prevents $\S_{a}$ from developing a non-trivial vacuum expectation value. A thorough analysis of the coupling between $ \S_{a} $ and chiral and anti-chiral scalar superfields indicate that no spontaneous breaking takes place. In components, the scalar $ \rho $ and the tensor $ \l_{\mu\nu} $ display a quartic coupling to the physical scalar components of the additional matter superfields, namely, $ \rho^{2} A_{-}^{2} $ and $ \l_{\mu\nu}^{2} A_{-}^{2} $, and no non-trivial minimum with $<A_{-}> \;\neq 0 $ shows up (a non-trivial minimum with non-zero vacuum expectation value restricted to the neutral scalar A is possible, but has no consequence for mass generation). So, spontaneous breaking does not happen to be a possibility for inducing a mass for the tensor $ \l_{\mu\nu} $. As a next step, the analysis of matter tensor fields in the framework of $ N=2 $ extended supersymmetries and the non-Abelian version of the $ N=1 $ model are to be pursued. \section{Conventions} \begin{equation} \sigma} \renewcommand{\S}{\Sigma^{\mu} = ( 1,\sigma} \renewcommand{\S}{\Sigma ), \hspace{.5cm} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu} = ( 1,-\sigma} \renewcommand{\S}{\Sigma ), \hspace{.5cm} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu}_{\dot{b}a} = \sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{b}}\;\; , \eqn{pauli} where $\sigma} \renewcommand{\S}{\Sigma = (\sigma} \renewcommand{\S}{\Sigma_{1},\sigma} \renewcommand{\S}{\Sigma_{2},\sigma} \renewcommand{\S}{\Sigma_{3})$ are the Pauli matrices \begin{equation} \sigma} \renewcommand{\S}{\Sigma_{1} = \left( \begin{array}{cc} 0 & 1 \\ 1 & 0 \end{array} \right) , \hspace{.5cm} \sigma} \renewcommand{\S}{\Sigma_{2} = \left( \begin{array}{cc} 0 & -i \\ i & 0 \end{array} \right), \hspace{.5cm} \sigma} \renewcommand{\S}{\Sigma_{3} = \left( \begin{array}{cc} 1 & 0 \\ 0 & -1 \end{array} \right) . \eqn{dirac-pauli} In addition, the matrices $\sigma} \renewcommand{\S}{\Sigma^{{\mu\nu}}$ and $\overline{\sigma} \renewcommand{\S}{\Sigma}^{{\mu\nu}}$ ( $ (\frac{1}{2},0)$ and $ (0,\frac{1}{2})$ SO(1,3) generators) are given by \begin{equation}\begin{array}{rl} \sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\nu \dot{a}b} &= \eta^{{\mu\nu}}\d_{a}^{\; b} - i (\sigma} \renewcommand{\S}{\Sigma^{{\mu\nu}})_{a}^{\; b} \\[2mm] \overline{\sigma} \renewcommand{\S}{\Sigma}^{\mu\dot{a} a}\sigma} \renewcommand{\S}{\Sigma^{\nu}_{a\dot{b}} &= \eta^{{\mu\nu}}\d^{\dot{a}}_{\; \dot{b}} - i (\overline{\sigma} \renewcommand{\S}{\Sigma}^{{\mu\nu}})^{\dot{a}}_{\; \dot{b}} \; , \end{array}\eqn{matrizes} and the trace is \begin{equation}\begin{array}{rl} \sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\sigma} \renewcommand{\S}{\Sigma}^{\nu\dot{a}b}\sigma} \renewcommand{\S}{\Sigma^{\a}_{b\dot{b}} \overline{\sigma} \renewcommand{\S}{\Sigma}^{\b\dot{b}a} = 2\displaystyle{\Biggl(} \eta^{{\mu\nu}}\eta^{\a\b} - \eta^{\mu\a}\eta^{\nu\b} + \eta^{\mu\b}\eta^{\nu\a} + i \varepsilon^{\mu\nu\a\b} \displaystyle{\Biggr)}, \end{array}\eqn{traco} where $\varepsilon^{0123} = -\varepsilon_{0123} = 1$. \\[2mm] The summation convention is: \begin{equation} \t\eta = \t^{a}\eta_{a}, \hspace{.5cm} \overline{\t}\overline{\eta} = \overline{\t}_{\dot{a}}\overline{\eta}^{\dot{a}}\; , \eqn{sobedesce} where lowering and raising of indices are effected through \begin{equation} \t^{a} = \varepsilon^{ab}\t_{b}, \hspace{.5cm} \t_{a} = \varepsilon_{ab}\t^{b}, \eqn{lorai} with $ \varepsilon_{ab} = - \varepsilon_{ba} $ , (the same for dotted indices). Differentation with respect to the anticommuting parameters $\t_{a}$ , $\overline{\t}_{\dot{a}}$ is defined by \begin{equation} \frac{\partial}{\partial\t^{a}}\t^{b} = \d_{a}^{\, b} \hspace{.5cm} \frac{\partial}{\partial\overline{\t}^{\dot{a}}}\overline{\t}^{\dot{b}} = \d_{\dot{a}}^{\, \dot{b}}. \eqn{delta1} Covariant derivatives with respect to the supersymmetry transformations are: \begin{equation}\begin{array}{rl} D_{a} & = \frac{\partial}{\partial\t^{a}} - i\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\overline{\t}^{\dot{a}}\partial_{\mu} \\[2mm] \overline{D}_{\dot{a}} & = - \frac{\partial}{\partial\overline{\t}^{\dot{a}}} + i \t^{a}\sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\partial_{\mu}, \end{array}\eqn{superderivadas} and they obey the anticommutation relations \begin{equation} \{ D_{a},\overline{D}_{\dot{a}} \} = 2 i \sigma} \renewcommand{\S}{\Sigma^{\mu}_{a\dot{a}}\partial_{\mu} \; ; \hspace{.5cm} \{D_{a},D_{b} \} = 0 = \{ \overline{D}_{a},\overline{D}_{\dot{b}} \}. \eqn{relations} The Dirac matrices, $ \gamma^{\mu} $ , playing a role in the purely physical spinorial kinetic term, $ i\;\overline\Psi \gamma^{\mu}\partial_{\mu} \Psi $ , have the following expression in terms of the Pauli matrices: \begin{equation} \gamma^{\mu} \; = \; \left( \begin{array}{cc} 0 & \sigma} \renewcommand{\S}{\Sigma^{\mu} \\ \overline\sigma} \renewcommand{\S}{\Sigma^{\mu} & 0 \end{array} \right) \; . \eqn{Gamma} The spinor $\overline\Psi $ is defined as usually: \begin{equation} \overline\Psi = \Psi^{\dagger}\gamma^{0}\; , \;\mbox{where} \;\;\gamma^{0} \; = \; \left( \begin{array}{cc} 0 & 1 \\ 1 & 0 \end{array} \right) \; . \eqn{Barrado} \vspace{2cm} \noindent{\large{\bf Acknowledgements}} We wish to thank Dr. M. A. de Andrade for helpful discussions on an earlier manuscript. The {\it Conselho Nacional de Desenvolvimento Cientifico e Tecnologico}, $CNPq$-Brasil is gratefully acknowledged for the financial support.
\section{Introduction} The strong CP and U(1) problems are two outstanding problems in the QCD sector of the standard model. The former consists in the gross disagreement of theoretical estimates of electric dipole moment of neutron (NEDM), which are invariably larger, by nearly nine to ten orders of magnitude, than the experimental upper limit [1]. The U(1) problem epitomises the difficulties in formulating a theoretically consistent framework to interpret the mass of the flavour singlet pseudoscalar meson $\eta^{'}$ [2], which, unlike the other Goldstone bosons, is very heavy. The genesis of both the problems is the anomaly for space-time indepedent `global' chiral rotation of fermi (quark) fields \begin{equation} q(x) \rightarrow e^{i\alpha\gamma_{5}} q(x), \hspace{.5cm} \bar{q}(x) \rightarrow \bar{q}(x) e^{i\alpha\gamma_{5}} \end{equation} In perturbation theory there is no trace of this `global' anomaly [3]. A space-time indepedent chiral phase in fermion mass drops out from all amplitudes diagram by diagram if the interactions are chirally invariant. This is not in conflict with the ABJ anomaly in the four divergence of the axial vector current which arises from the triangle diagram in perturbation theory. The `global' anomaly is just the space-time integral of the ABJ anomaly, which, in path integral approach, corresponds to space-time dependent `local' chiral transformations \begin{equation} q(x) \rightarrow e^{i\alpha(x)\gamma_{5}} q(x), \hspace{.5in} \bar{q}(x) \rightarrow \bar{q}(x) e^{i\alpha(x)\gamma_{5}} \end{equation} In perturbation theory the integrand is nontrivial but the integral vanishes. The sine qua non of a nonvanishing integral and hence of `global' chiral anomaly is the zero modes of Euclidean Dirac operator which live in compactified space-time and are inaccessible in perturbative framework. The carrier of the virus of fermion zero modes in popular path integral approach is identified to be the Osterwalder-Schrader (OS) [4] recipe for the conjugate Dirac field in Euclidean metric, viz., $\bar{q}(x)$ is indepedent of $q(x)$. In relativistic metric the relation $\bar{q}(x) = (\gamma_{o} q(x))^{+}$ relates the Dirac field to its conjugate. The OS recipe, therefore, requires that the degrees of freedom are doubled in Euclidean metric. This is neither natural nor necessary. We propose a prescription which is just the opposite of the OS recipe. To be precise, we require $\bar{q}(x)$ to be antilinear in $q(x)$ and the relationship to obey reciprocity [5]. The representation of $\bar{q}(x)$ is unique if it is required that $\bar{q}(x)$ has the correct chiral properties, i.e., obeys eq.(1). The novel representation reproduces the correct two-point correlation function, and hence, by antisymmetry, all the 2n-point fermion correlation functions of perturbation theory [5].This assures that the novel representation is consistent with all the standard axioms of Euclidean field theory [6]. The point of interest is that the fermion zero modes are evaded and the novel representation leads to a formulation of QCD which is free from `global' chiral anomaly. This solves the strong CP and the U(1) problems [5]. It is remarkable that the novel representation yields, in path integral approach, a nonvanishing determinent of the Dirac operator $D\!\!\!\!/\,$ in the nontrivial sector $\nu \not= 0$ of gauge fields \begin{equation} \nu \equiv \frac{g^{2}}{16\pi^{2}} \int tr F_{\mu\nu} \tilde{F}_{\mu\nu} d^{4}x \end{equation} whereas, the OS recipe gives a vanishing result. This reflects an inherent ambiguity in the definition of the determinant of Dirac operator in non-trivial sector. The ambiguity becomes transparent in Weyl space where one has the option to write Dirac determinant either as \begin{equation} {\em det} D\!\!\!\!/\, = {\em det} (-DD^{+}) {\em or}\hspace{.1in} {\em det} (-D^{+}D) \end{equation} in terms of Weyl operators $D, D^{+}$ defined through \begin{eqnarray} D\!\!\!\!/\, &\equiv &\gamma_{\mu} (i\partial_{\mu} - g A_{\mu}) \nonumber \\ &= &(\begin{array}{cc} 0 &D\\ D^{+} &0\end{array}) \end{eqnarray} In nontrivial sector $\nu \not= 0$ one of the options in (4) vanishes while the other does not. This follows from the index theorem \begin{eqnarray} \nu &= &{\em dim} {\em ker} (DD^{+}) - {\em dim} {\em ker} (D^{+}D) \nonumber \\ &= &n_{+} - n_{-} \end{eqnarray} and the theorem that there are no `wrong chirality' solutions of Dirac operator [7], i.e., for $\nu \geq 0$ $(\nu \leq 0)$, the number of normalisable negative (positive) chirality solution $n_{-} (n_{+})$ is zero. The novel representation chooses the nonvanishing option for the Dirac determinant while the OS recipe leads to the vanishing option in nontrivial sector. Resolution of the strong CP and the U(1) problems, and hence, it appears, physics chooses the option of the novel representation. \section{Global Chiral Anomaly} In Euclidean metric the QCD action is given by \begin{equation} S_{QCD} = S_{G} + S_{F} + \theta_{ew} \Delta^{'}S + \gamma_{QCD} \Delta S \end{equation} where $S_{F}$ is the fermionic piece \begin{equation} S_{F} \equiv \int \bar{q}(x) (D\!\!\!\!/\, - iM) q(x) d^{4}x \end{equation} and $S_{G}$ is the gluon action. The (light) quarks have three flavours. For convenience, we assume the mass matrix M to be diagonal in flavour space and suppress the flavour indices. The two terms $\Delta S$ and $\Delta^{'}S$ are potential sources of CP violation. While the $\theta_{QCD}$ term is attributed to the topological structure of the QCD vacuum \begin{equation} \Delta S \equiv \frac{g^{2}}{16\pi^{2}} \int tr F_{\mu\nu} \bar{F}_{\mu\nu} d^{4}x \end{equation} the chiral phase $\theta_{ew}$ in quark mass \begin{equation} \Delta^{'}S \equiv \int \bar{q}(x) M\gamma_{5} q(x) d^{4}x \end{equation} arises from Higgs interactions in the electroweak sector. In compactified space-time $\Delta S$ assumes integral values $\nu$. The degrees of freedom of $q(x)$ are the Grassmann gennators which appear as coefficients in the expansion of $q(x)$ in a complete set of basis functions.For convenience, the orthonormal set of eigenfunctions of the Dirac operator are chosen as basis functions, \begin{equation} q(x) = \sum_{r} (a_{r} + a_{-r} \gamma_{5}) \varphi_{r}(x) + \sum_{i} a_{oi} \varphi_{oi}(x) \end{equation} where the normalised eigenfunctions obey the equations, \begin{eqnarray} D\!\!\!\!/\, \varphi_{r}(x) = \lambda_{r} \varphi_{r} (x) &, &\hspace{.5in} D\!\!\!\!/\, \gamma_{5}\varphi_{r}(x) = - \lambda_{r}\gamma_{5}\varphi_{r}(x) \nonumber \\ D\!\!\!\!/\, \varphi_{oi}(x) = 0 &, &\hspace{.5in} \gamma_{5}\varphi_{oi}(x) = \epsilon_{i}\varphi_{oi} \end{eqnarray} The zero eigenmodes $\varphi_{oi}$ have definite chiralities, positive $(\epsilon_{i} = 1)$ or negative $(\epsilon_{i} = -1)$. The OS recipe is implemented by choosing an independent set of Grassmann generators for $\bar{q}(x)$ \begin{equation} \bar{q}(x) = \sum (b_{r} + b_{-r} \gamma_{5}) \varphi_{r}(x) + \sum b_{oi}\varphi_{oi}(x) \end{equation} The Jacobian of the measure in the fermionic partition function $Z_{F}$ \begin{equation} Z_{F} \equiv \int d\mu {\em exp} [- S_{F}] \end{equation} has the form \begin{equation} ln J(\alpha) = - 2i\alpha \int A(x)d^{4}x \end{equation} for `global' chiral rotation (1). The integrand A(x) is identified as the `local' chiral anomaly. The measure corresponding to OS recipe for $\bar{q}(x)$ is given by $$d\mu^{I} = \Pi_{r} da_{r} db_{r} \Pi da_{oi} db_{oi}$$ The `local' chiral anomaly for this measure was obtained by Fujikawa [8], \begin{eqnarray} A^{I}(x) &= &2\sum\varphi_{r}^{+}(x)\gamma_{5}\varphi_{r}(x)+\sum\epsilon_{i}\varphi^{+}_{oi}(x)\varphi_{oi}(x)\nonumber\\ &= &\frac{g^{2}}{16\pi^{2}} trF_{\mu\nu}(x) \bar{F}_{\mu\nu}(x) \end{eqnarray} Nonzero eigenmodes drop out because of orthogonality of $\varphi_{r}(x)$ and $\gamma_{5}\varphi_{r}(x)$ and only zero modes of $A^{I}(x)$ survive in the integral (15) for the Jacobian \begin{eqnarray} ln J^{I}(\alpha) &= &-2i\alpha(n_{+} - n_{-}) \nonumber \\ &= &-2i\alpha\frac{g^{2}}{16\pi^{2}} \int tr F_{\mu\nu}\bar{F}_{\mu\nu}d^{4}x \end{eqnarray} This means that under `global' chiral rotation QCD action changes according to the formula \begin{equation} S_{QCD} \rightarrow S^{I}_{QCD}(\alpha) = S_{G} + S_{F} + (\theta_{ew} + 2\alpha) \Delta^{'}S + (\theta_{QCD} - 6\alpha)\Delta S \end{equation} In effective Lagrangians for chiral models there is no scope for a nontrivial Jacobian. The global chiral U(1) anomaly (17) in underlying QCD can, therefore, be reproduced in effective Lagrangians through an `anomaly term' which breaks chiral symmetry explicitly. A popular representation of the anomaly term is [9] \begin{equation} \Delta S^{I}_{eff} = - m^{2}_{\eta^{'}} f^{2}_{\pi} \int [tr ln(\frac{U}{U^{+}}) - \theta_{QCD}]^{2}d^{4}x \end{equation} where $f_{\pi}$ and $m^{'}_{\eta}$ are respectively the pion decay constant and the mass of the flavour singlet Goldstone boson. The meson matrix U transforms as $U \rightarrow Ue^{i\alpha}$ under global chiral rotation (1). The problems with this `anomaly term' are (a) its chiral variation depends explicitly on $\theta_{QCD}$, and (b) its second order variation does not vanish. Neither of these properties hold in the underlying QCD. This is the crux of the controversy between 't Hooft and Crewther [2], and the reason why the popular resolution (19) of the U(1) problem is regarded as unsatisfactory. The transformation law (18) shows that neither $\theta_{ew}$ nor $\theta_{QCD}$ can be physical. Only the chirally invariant combination $\bar{\theta} \equiv (\theta_{QCD} + 3\theta_{ew})$ can appear in CP violating effects. Theoretical estimates in various chiral models suggests [1] NEDM in the range $$d_{n} \approx \bar{\theta} \times 10^{-15 \pm 1} e.cm$$ Experimental upper limit $\vert d_{n}\vert < 10^{-25} e.cm$, therefore, puts a stringent constraint $\bar{\theta} < 10^{-10}$. This is the crux of the strong CP problems. The strong CP problem is, therefore, a serious problem of fine tuning. Two parameters $\theta_{QCD}$ and $3\theta_{ew}$ which are of completely different origins in the standard model must be so fine tuned as to cancel each other completely. \section{Novel representation of Euclidean Dirac fermion} We start from the ansatz which is just the opposite of the OS recipe, i.e., we assume that in Euclidean metric the conjugate field $\chi(x) \equiv [\bar{q}(x)]^{+}$ is linear in $q(x)$. This means that the Grassmann generators defining the degrees of freedom of $\chi(x)$ are a subset of the generators $\{a_{r}, a_{oi}\}$ appearingin $q(x)$. The resulting representation of $\chi(x)$ is unique, modulo an overall sign, if one requires that, (a) chiral charge of $\chi(x)$ is opposite to that of $q(x)$, i.e., if $q(x) \rightarrow e^{i\alpha\gamma_{5}} q(x)$ then $\chi(x) \rightarrow e^{-i\alpha\gamma_{5}} q(x)$, and (b) the linear relation obeys reciprocity, \begin{equation} \chi(x) = \sum_{r}[a_{r} - a_{-r} \gamma_{5}]\varphi_{r}(x) \end{equation} The crucial point to note is that $\chi(x)$ cannot contain the zero mode generators $a_{oi}$ which transform as $a_{oi} \rightarrow a_{oi} e^{i\epsilon_{i}\alpha}$ under chiral rotation (1). This will be in conflict with the chiral charge of $\chi(x)$. As a result, the fermion action $S_{F}$ is devoid of the zero mode generators \begin{eqnarray} S_{F} &= &\int \chi^{+}(x) (D\!\!\!\!/\, - iM) q(x)d^{4}x \nonumber \\ &= &\sum_{r} [\lambda_{r}(a^{*}_{r}a_{r} + a^{*}_{-r}a_{-r}) - iM (a^{*}_{r}a_{r} - a^{*}_{-r}a_{-r})] \end{eqnarray} The measure appropriate for this action $$d\mu = \Pi_{r} da^{*}_{r} da_{r} da^{*}_{-r} da_{-r}$$ leads to the partition function \begin{equation} Z_{F} = \Pi_{\lambda_{r}>o} (\lambda_{r}^{2} + M^{2}) \end{equation} whose chiral limit (M=0) does not vanish in the nontrivial sector $(\nu \not= 0)$ of gauge fields. The two-point correlation function, obtained in the usual path integral approach, coincides with the familiar formula \begin{equation} < q(x) \bar{q}(y) > = < x\vert\frac{1}{D\!\!\!\!/\, - iM}\vert y > - \frac{\sum_{i}\varphi_{oi}(x)\varphi^{+}_{oi}(y)}{-iM} \end{equation} except that the zero mode contributions are subtracted out. In the limit of zero coupling $g = 0$, the momentum representation of the correlation function coincides with the Wick-rotated relativistic Feynman propagator \begin{equation} < q(x) \bar{q}(y) >_{g=0} = \frac{1}{(2\pi)^{4}} \int d^{4}p \frac{p\!\!\!/ + iM}{p^{2} + M^{2}} e^{-ip(x-y)} \end{equation} The remaining 2n-point correlation functions follow, in path integral framework, from the anticommutation of $q(x)$ and $\bar{q}(y)$ \begin{equation} < q(x_{1}) ... q(x_{m}) \bar{q}(y_{1}) ... \bar{q}(y_{n}) > = \delta_{mn} {\em det} [< q(x_{i}) \bar{q}(y_{i})] \end{equation} The `local' chiral anomaly is given by an expression analogous to that in the OS formulation (16) except that zero modes are excluded from the sum on the right hand side, \begin{eqnarray} A^{II(x)} &= &2\sum_{\lambda_{r}>0} \varphi^{+}_{r}(x) \gamma_{5} \varphi_{r}(x) \nonumber \\ &= &\frac{g^{2}}{16\pi^{2}} \int tr F_{\mu\nu} \bar{F}_{\mu\nu} d^{4}x - \epsilon_{i}\varphi^{+}_{oi}(x) \varphi_{oi}(x) \end{eqnarray} Thus the four divergence of the U(1) axial current has a nontrivial anomaly $A^{II}(x)$ \begin{equation} \partial_{\mu} J_{\mu 5} (x) = 2\bar{q}(x) M \gamma_{5} q(x) + A^{II}(x) \end{equation} which coincides, as it must, with the perturbative (absence of zero modes) ABJ anomaly. However, the `global' chiral anomaly vanishes \begin{eqnarray} ln J^{II}(\alpha) &= &- 2i\alpha \int A^{II}(x) d^{4}x \nonumber \\ &= &0 \end{eqnarray} and our desired goal is achieved. The vanishing is the direct consequence of the orthogonality of nonzero eigenmodes $\varphi_{r}(x)$ and $\gamma_{5}\varphi_{r}(x)$, or, if one prefers, the index theorem. The novel representation (20) is equivalent to the functional relation \begin{equation} \chi(x) = \frac{1}{[D\!\!\!\!/\, ^{2}]^{\frac{1}{2}}}D\!\!\!\!/\, q(x) \end{equation} which is nonlocal. This is not a disability in Euclidean field theory. Locality is a field theoretic axiom only in rleativistic metric. This translates in Euclidean field theory into the axiom of (anti) symmetry of correlation functions under permutations [6]. The fact that all the correlation functions are reproduced correctly through the eqs.(24) and (25) assures us that the novel representation (20, 29) is consistent not only with the axiom of symmetry but with the other axioms of Euclidean field theory, e.g., reflection positivity, cluster decomposition etc., as well. \section{Resolution of strong CP and U(1) problems} In the scenario corresponding to the novel representation (20, 29) the QCD action is invariant under `global' chiral rotation in the chiral (M = 0) limit \begin{equation} S_{QCD} \rightarrow S^{II}_{QCD}(\alpha) = S_{G} + S_{F} + (\theta_{ew} + 2\alpha) Delta^{'}S + \theta_{QCD} \Delta S \end{equation} In effective Lagrangians, the `anomaly' term, which is invariant under `global' chiral rotation but reproduces the ABJ anomaly in axial Ward identity, is easily constructed \begin{equation} \Delta S^{II}_{eff} = - m^{2}_{\eta^{'}} f^{2}_{\pi} \int [tr ln (\frac{U}{U^{+}}) - < tr ln (\frac{U}{U^{+}})>]^{2} d^{4}x \end{equation} The controversial features of the popular construction (19) thus disappear and the U(1) problem satisfactorily resolved [5]. The law of transformation (30) shows that the parameter $\theta_{QCD}$, the coefficient of $\Delta S$, is invariant, while $\theta_{ew}$ is unphysical and can be eliminated trivially by `global' chiral rotation (1). There is no longer any problem of fine tuning and CP symmetry of QCD is simply ensured through the `natural' choice $\theta_{QCD} = 0$ [5]. We conclude that the strong CP and the U(1) problems both are legacies of the OS recipe. The alternative scenario of QCD with the novel representation (20, 29) for the conjugate Dirac field is not afflicted with these blemishes [5].
\section{\bf 1. Introduction} The theory of spinors in spaces with boundaries is of interest physically in connection with quantum cosmology and supergravity. (See D'Eath and Esposito [\putref{DandE}] and Esposito [\putref{Esposito}] for some history of these questions.) In mathematics it is encountered in the spin-index theorem and the Atiyah, Patodi and Singer $\eta$ spectral asymmetry function, the standard reference being Gilkey's book, [\putref{gilk}]. As explained in [\putref{DandE}], for self-adjointness of the Dirac operator, there is a choice between spectral and local (mixed) boundary conditions, the former being of relevance for the spin-index and the latter having more physical significance in connection with supersymmetry, string theory and quantum gravity, [\putref{MandP2,Luck}], although in the guise of relative conditions they do have a cohomological importance, [\putref{gilk,BandG2}]. In the special case of the Euclidean 4-ball, it was shown [\putref{DandE2,Kam, KandM}] that the value of $\zeta(0)$, which determines the scaling of the theory, was the same for both sets of conditions. In this note we report on the same question for the one-loop effective action, which is, up to factors, $\zeta'(0)$. Our method will be that explained in [\putref{Dow9}]. \section{\bf 2. Mode properties and calculation} The analysis of the modes of the massless Dirac equation on the 4-ball was carried out by D'Eath and Esposito [\putref{DandE,DandE2}] and we will do no more here on this matter than use their results. For local boundary conditions they found that the eigenvalues, $\alpha^2$, are the roots of the equation $$ F^L_p(\alpha)=J_{p-1}^2(\alpha)-J_p^2(\alpha)=0 \eql{locb}$$with a degeneracy, for a given $p$, of $p^2-p$, $p=1,2,\ldots$. For spectral conditions, there is the simpler, scalar-like condition, $$ F_p^S(\alpha)=J_p(\alpha)=0 \eql{specb}$$ with degeneracy $2(p^2+p)$, $p=1,2,\ldots$. Our approach is based on the Mittag-Leffler decomposition, $$ z^{-\beta}F_p(z)=\gamma\prod_\alpha\bigg(1-{z^2\over\alpha^2}\bigg), \eql{ml1}$$ where $$\eqalign{ &\beta=p,\quad\gamma={1\over2^pp!},\quad{\rm spectral},\cr &\beta=2(p-1),\quad\gamma={1\over\big(2^{p-1}(p-1)!\big)^2},\quad{\rm local}. \cr} $$ This standard decomposition was earlier employed by Moss [\putref{Moss}] and by D'Eath and Esposito [\putref{DandE}] when looking at the heat-kernel expansion and $\zeta(0)$. Here, when finding $\zeta'(0)$, we need the normalising factor, $\gamma$, which follows from the small-$z$ behaviour of $F_p(z)$. A few details of the calculation will be given but, for brevity, some of our previous work must be utilised. Bypassing a number of steps, which are fully explained in [\putref{Dow9,Dow8}], we define the quantities $$ G_N\sim\sum_{p=1}^\infty p^N\bigg[\big(p-{1\over2}\big) \ln{2p\over p+\epsilon}+(\epsilon-p)+\sum_{n=1}^{N+1}\bigg({E_n(t)\over \epsilon^n} -{E_n(1)\over p^n}\bigg)+I_N(p)\bigg] \eql{logdetl}$$ and $$ H_N\sim\sum_{p=1}^\infty p^N\bigg[p\ln{2p \over p+\epsilon}+\epsilon-p-{1\over2}\ln{\epsilon\over p}+\sum_{n=1}^{N+1} \bigg({T_n(t)\over \epsilon^n}-{T_n(1)\over p^n}\bigg)+I_N(p)\bigg], \eql{logdets}$$ with $$ I_N(p)=\int_0^\infty\bigg({1\over2} -{1\over\tau}+\sum_{k=1}^{[N/2]+1}(-1)^kB_{2k}{\tau^{2k-1}\over(2k)!} +{1\over e^\tau-1}\bigg){e^{-\tau p}\over\tau}\,d\tau, $$ in terms of which we can write the spin-half quantities, $$\eqalign{ &{\zeta_{1/2}^L}'(0)=2\big(G_2-G_1\big),\cr &{\zeta^S_{1/2}}'(0)=2\big(H_2+H_1\big).\cr} \eql{zedashes}$$ The labels $S$ and $L$ refer to spectral and local boundary conditions respectively. In equations (\puteqn{logdetl}) and (\puteqn{logdets}) the $\sim$ symbol signifies that the mass-independent part of the large-mass asymptotic limit is to be taken. The $E_n(t)$ are the polynomials in $t=p/\epsilon$, $\epsilon=(m^2+p^2)^{1/2}$, that occur in the asymptotic expansion of $F_p^L(im)$ of (\puteqn{locb}) derived by D'Eath and Esposito (they call them $A_n/2$) from Olver's series. The $T_n(t)$ are the corresponding polynomials for the scalar case, [\putref{Moss,Dow8}]. The condition that makes equation (\puteqn{logdetl}) possible is $E_n(1)=T_n(1)$ which can be proved from the explicit definition of the $E_n$. We note that $T_n(1)$ is zero for $n$ even and that $T_{2k-1}(1)=(-1)^kB_{2k}/2k(2k-1)$ in terms of Bernoulli numbers. We have made use of the algebraic results of D'Eath and Esposito, [\putref{DandE}] section IV, in deriving (\puteqn{logdetl}). Expression (\puteqn{logdets}) is identical to one occurring for scalar fields on the even ball, except that $N$, there being the power of $p$ in the expansion of the degeneracy, is even. Hence for $N=2$, our previous result in [\putref{Dow8,DandA2}] for the 4-ball (see also [\putref{BGKE}]) could be used without change. {}From the technique outlined in [\putref{Dow9}] the following useful limits can be deduced,\marginnote{work out ln m part} $$\eqalign{ &\sum_{p=1}^\infty p^N(\epsilon-p)\sim -\zeta_R(-N-1)+O(\ln m),\cr &\sum_{p=1}^\infty p^N\ln\big({2p\over p+\epsilon}\big)\sim-\zeta_R'(-N)+\ln2\, \zeta_R(-N)+O(\ln m),\cr &\sum_{p=1}^\infty p^N\ln\big({\epsilon\over p}\big)\sim\zeta_R'(-N)+O(\ln m).\cr} \eql{aslims}$$ It is necessary to state that a hidden regularisation has been employed to render the summations finite. This consists of removing sufficient of the Taylor expansion of the summand and will not be indicated. Since the entire expression is finite, the divergent terms so introduced must all cancel. These limits enable some of the terms in (\puteqn{logdetl}) and (\puteqn{logdets}) to be dealt with quickly. The rest, {\it i.e. } the polynomial and integral contributions, need a little more work. We write them as in [\putref{Dow8,Dow9}], $$\eqalign{ \sum_{p=1}^\infty&\, p^N\bigg[\sum_{n=1,3,\ldots}^{N+1} P_n(1) \bigg({1\over\epsilon^n}-{1\over p^n}\bigg) +\sum_{n=1}^{N+1}{P'_n(t)\over\epsilon^n}\bigg]\cr &+\lim_{s\to0}\!\int_0^\infty\!\!\bigg({1\over2}\! -\!{1\over\tau}\!+\!\sum_{k=1}^{[N/2]+1}\!\!(-1)^kB_{2k}{\tau^{2k-1} \over(2k)!}\! +\!{1\over e^\tau-1}\bigg)\tau^{s-1}(-1)^N{d^N\over d\tau^N} {1\over e^\tau\!-\!1}d\tau,\cr} \eql{rem2}$$ where $P_n$ stands for either $E_n$ or $T_n$ and $P'_n(t)=P_n(t) -P_n(1)$. A recursion is developed for the multiple derivative in (\puteqn{rem2}) and the contribution from the integral found to be, after some algebra, $$ \zeta_R'(-N-1)+{1\over2}\zeta_R'(-N)+\zeta_R(-N-1)+ \sum_{k=1}^{N+1}M_k^{(N)}\zeta_R'(-k), \eql{intcont}$$ where the coefficient matrix $M$ is defined by $$ M_k^{(N)}=\sum_{l=k}^{N+1}A_l^{(N)}{S_{l+1}^{(k+1)}\over l!} $$ in terms of easily evaluated recursion constants $A_l^{(j)}$ and Stirling numbers $S_l^{(k)}$, [\putref{Dow8,Dow9}]. Assembling the various pieces, and using special values for the $M_k^{(N)}$, we find $$\eqalign{ G_N&={\zeta_R'(-N)\over2}\!+\!{\zeta_R'(-N-1)\over N+1}\! +\!\sum_{k=1}^{N-1}\!M_k^{(N)}\zeta_R'(-k)\!\cr &+\!\big({1\over2}\zeta_R(-N)-\zeta_R(\!-\!N\!-1)\big)\,\ln2 +\int_0^1t^NE''_{N+1}(t)\,dt+L_N,\quad N\ge1,\cr} \eql{totl}$$ where $P_n''(t)=P'_n(t)/(1-t^2)$ and $$ L_N=T_{N+1}(1)\bigg(\ln2 +\sum_{k=1}^N{1\over k}+\sum_{q=1}^{N/2} {(-1)^q\sqrt\pi\,(N/2)!\over2q(N/2-q)!\Gamma(q+1/2)}\bigg). $$ The last two terms in (\puteqn{totl}) come from the first line of (\puteqn{rem2}). Explicitly for $N=0$, a case needed later, $$ G_0=-{1\over24}+{1\over12}\ln2+\zeta_R'(-1). \eql{gee0}$$ For spectral conditions, $$\eqalign{ H_N&=-{\zeta_R'(-N)\over2}+{\zeta_R'(-N-1)\over N+1} +\sum_{k=1}^{N-1}M_k^{(N)}\zeta_R'(-k)\cr &+\zeta_R(-N-1)\,\ln2 +\int_0^1t^NT''_{N+1}(t)\,dt+L_N,\quad N\ge1,\cr} \eql{tots}$$ and $$ H_0= {5\over24}-{1\over6}\ln2+\zeta_R'(-1)-\zeta_R'(0). \eql{aitch0}$$ Making the constructions (\puteqn{zedashes}), one finds for local spin-half, \marginnote{SPINHALF.MTH} $$ {\zeta^L_{1/2}}'(0)={251\over15120}-{11\over180}\ln2 +{2\over3}\big(\zeta_R'(-3)-\zeta_R'(-1)\big)\approx0.088108 \eql{dashl}$$ and for spectral, $$ {\zeta^S_{1/2}}'(0)=-{2489\over30240}+{1\over45}\ln2 +{2\over3}\big(\zeta_R'(-3)-\zeta_R'(-1)\big)\approx0.046962 \eql{dashs}$$ which are the main results of this note. The specific forms of the $E_n$ polynomials given in [\putref{DandE}], have been used to evaluate the integrals in (\puteqn{totl}). We remark that in the corresponding evaluation of $\zeta(0)$ (=$11/360$), one needs only the particular value $P_N(1)$, which equals $\zeta_R(-N)/N$, a non-transcendental, local quantity. \section{\bf 3. Higher spins} The eigenvalue conditions for some higher-spin theories are summarised in [\putref{DandE2}] section VI. A mechanical application of the present technique yields the following results. For real spin-0 with Dirichlet conditions, $$\eqalign{ 2{\zeta^D_0}'(0)&=2H_2\cr \noalign{\vskip5truept} &={173\over15120}+{1\over45}\ln2+{2\over3}\zeta_R'(-3)-\zeta_R'(-2) +{1\over3}\zeta_R'(-1)\cr \noalign{\vskip5truept} &\approx0.005738.\cr} $$ For spin-1 (Maxwell) with Dirichlet (magnetic) conditions, $$\eqalign{ \zeta'_{\rm TV}(0)&=2(H_2-2H_0)\cr \noalign{\vskip5truept} &=-{6127\over15120}+{16\over45}\ln2+{2\over3} \zeta_R'(-3)-\zeta_R'(-2)-{5\over3}\zeta_R'(-1)+2\zeta_R'(0)\cr \noalign{\vskip5truept} &\approx-1.68691.\cr} \eql{TV}$$ For spin-3/2 physical degrees of freedom with spectral conditions, $$\eqalign{ {\zeta^S_{3/2}}'(0)&=2(H_2+H_1-H_0)\cr \noalign{\vskip5truept} &=-{27689\over30240}+{31\over45}\ln2+{2\over3}\zeta_R'(-3) -{14\over3}\zeta_R'(-1)+4\zeta_R'(0)\cr \noalign{\vskip5truept} &\approx-3.33834.\cr} \eql{3/2s}$$ These results imply, rather trivially, the sum rule, $$ {\zeta^S_{3/2}}'(0)-{\zeta^S_{1/2}}'(0)=2\big(\zeta_{\rm TV}'(0) -2{\zeta^D_0}'(0)\big). \eql{ident1}$$ The same relation holds also for $\zeta(0)$, $$ {\zeta^S_{3/2}}(0)-{\zeta^S_{1/2}}(0)=2\big(\zeta_{\rm TV}(0) -2\zeta^D_0(0)\big), \eql{ident2}$$ and, in fact, for all coefficients in the heat-kernel expansion, as can be checked numerically from the tables provided in [\putref{BEK}] and [\putref{KandC}]. The specific values, $$ \zeta^S_{3/2}(0)=-{289\over360},\quad\zeta^S_{1/2}(0)={11\over360}, \quad\zeta_{\rm TV}(0)=-{77\over180},\quad \zeta^D_0(0)=-{1\over180}, \eql{zezeros} $$ were computed in references [\putref{DandE,DandE2,Louko}], see also [\putref{MandP,KandC,Poletti}]. The spectral label, $S$, can be replaced by the local one, $L$, in (\puteqn{zezeros}). The sum rules are only special cases of the general relation $$ {\zeta^S_{3/2}}(s)-{\zeta^S_{1/2}}(s)=2\big(\zeta_{\rm TV}(s) -2\zeta^D_0(s)\big), \eql{ident3}$$ which is a consequence of the eigenvalue condition, (\puteqn{specb}), and the various quadratic degeneracies. For spin-2 transverse-traceless modes with Dirichlet conditions, [\putref{Schleich}], \ {\it i.e. } $$ F_p^{\rm TT}=J_p(\alpha)=0 $$ and degeneracy $2(p^2-4)$, $p\ge3$, we find $$\eqalign{ \zeta'_{\rm TT}(0)&=2(\overline H_2-4\overline H_0)=2(H_2-H_0)+ 6\big(\zeta_R'(0)+\ln2\big)\cr &= -{25027\over15120}+{331\over45}\ln2+{2\over3}\zeta_R'(-3)-\zeta_R'(-2) -{23\over3}\zeta_R'(-1)+14\zeta_R'(0)\cr \noalign{\vskip5truept} &\approx-8.119619,\cr} \eql{TT}$$ where the bar signifies that the $p=1$ term has been left out in (\puteqn{logdets}). (The easiest way of doing this is to remove the overall $p=1$ term at the outset.) For the record, the local spin-3/2 expression is $$\eqalign{ {\zeta^L_{3/2}}'(0)&=2(\overline G_2-\overline G_1-2\overline G_0)\cr &={2771\over15120}+{289\over180}\ln2+{2\over3}\zeta_R'(-3) -{14\over3}\zeta_R'(-1)+4\zeta_R'(0)\cr \noalign{\vskip5truept} &\approx-1.60405,\cr} \eql{3/2l}$$ which exhibits the anomaly value of $-289/360$. Arbitrary-spin fields can be treated in exactly the same way, most easily using the mode analysis given in [\putref{Dow10,DandC2}], and will be discussed in a later communication. \begin{ignore} For massless spin-j fields, the degeneracy is $2(n^2-j^2)$, [\putref{Dow10}], where $n=p$ for integer spins with Dirichlet conditions and $n=p+j$ for half odd-integer ones with spectral conditions. \end{ignore} \section{\bf 4. Comments} The above expressions for the $\zeta'(0)$ have also been obtained by Kirsten and Cognola [\putref{KandC}] using the method of Bordag {\it et al}, [\putref{BGKE}]. The local result, (\puteqn{dashl}), agrees with that of Apps, reported in [\putref{DandA2}] and found using a conformal transformation from the 4-hemisphere. In fact, the final expression in (\puteqn{dashl}) is $\zeta'_S(0)$ on the hemisphere, the rest coming from the cocycle function obtained from an integration of the conformal anomaly, as in [\putref{DandA,BandG}] for example. Spectral conditions are also conformally invariant and it seems that (\puteqn{dashs}) can be interpreted in a similar way. The same structure is also apparent in (\puteqn{3/2s}) and (\puteqn{3/2l}) for spin 3/2. This suggests that the eigenvalue problem on the hemisphere is the same, or is equivalent, for spectral and local boundary conditions. This is confirmed by, and may explain, the equality of $\zeta(0)$ for these conditions found by D'Eath and Esposito in flat space and by Kamenshchik and Mishakov on the bounded sphere. To the author's knowledge, the cocycle function has not been calculated for spectral conditions. The extension to higher, even-dimensional spaces is straightforward and simply consists of substituting (\puteqn{totl}) or (\puteqn{tots}) into the appropriate polynomial form of the spinor degeneracy. For odd dimensions the major difference is that the $p$-sums run over half odd-integers and presents no problem [\putref{Dow9}]. For example, the Maxwell modes on the 3-ball are classic, {\it e.g. } [\putref{deB}], and it is soon shown that the magnetic determinant is obtained by doubling the scalar Dirichlet value and subtracting $-2\ln2$ to allow for the different starting point of the mode sum. Similarly, the electric determinant is the double the scalar Robin one, with $\beta=1/2$, again minus $-2\ln2$. \section{\bf Acknowlegment} I wish to thank Klaus Kirsten for helpfully communicating his results. \vskip 10truept \noindent{\bf{References}} \vskip 5truept \begin{putreferences} \reference{Rayleigh}{Lord Rayleigh{\it Theory of Sound} vols.I and II, MacMillan, London, 1877,78.} \reference{KCD}{G.Kennedy, R.Critchley and J.S.Dowker \aop{125}{80}{346}.} \reference{Donnelly} {H.Donnelly \ma{224}{1976}161.} \reference{Fur2}{D.V.Fursaev {\sl Spectral geometry and one-loop divergences on manifolds with conical singularities}, JINR preprint DSF-13/94, hep-th/9405143.} \reference{HandE}{S.W.Hawking and G.F.R.Ellis {\sl The large scale structure of space-time} Cambridge University Press, 1973.} \reference{DandK}{J.S.Dowker and G.Kennedy \jpa{11}{78}{895}.} \reference{ChandD}{Peter Chang and J.S.Dowker \np{395}{93}{407}.} \reference{FandM}{D.V.Fursaev and G.Miele \pr{D49}{94}{987}.} \reference{Dowkerccs}{J.S.Dowker \cqg{4}{87}{L157}.} \reference{BandH}{J.Br\"uning and E.Heintze \dmj{51}{84}{959}.} \reference{Cheeger}{J.Cheeger \jdg{18}{83}{575}.} \reference{SandW}{K.Stewartson and R.T.Waechter \pcps{69}{71}{353}.} \reference{CandJ}{H.S.Carslaw and J.C.Jaeger {\it The conduction of heat in solids} Oxford, The Clarendon Press, 1959.} \reference{BandH}{H.P.Baltes and E.M.Hilf {\it Spectra of finite systems}.} \reference{Epstein}{P.Epstein \ma{56}{1903}{615}.} \reference{Kennedy1}{G.Kennedy \pr{D23}{81}{2884}.} \reference{Kennedy2}{G.Kennedy PhD thesis, Manchester (1978).} \reference{Kennedy3}{G.Kennedy \jpa{11}{78}{L173}.} \reference{Luscher}{M.L\"uscher, K.Symanzik and P.Weiss \np {173}{80}{365}.} \reference{Polyakov}{A.M.Polyakov \pl {103}{81}{207}.} \reference{Bukhb}{L.Bukhbinder, V.P.Gusynin and P.I.Fomin {\it Sov. J. Nucl. Phys.} {\bf 44} (1986) 534.} \reference{Alvarez}{O.Alvarez \np {216}{83}{125}.} \reference{DandS}{J.S.Dowker and J.P.Schofield \jmp{31}{90}{808}.} \reference{Dow1}{J.S.Dowker \cmp{162}{94}{633}.} \reference{Dow2}{J.S.Dowker \cqg{11}{94}{557}.} \reference{Dow3}{J.S.Dowker \jmp{35}{94}{4989}; erratum {\it ibid}, Feb.1995.} \reference{Dow5}{J.S.Dowker {\it Heat-kernels and polytopes} To be published} \reference{Dow6}{J.S.Dowker \pr{D50}{94}{6369}.} \reference{Dow7}{J.S.Dowker \pr{D39}{89}{1235}.} \reference{Dow8}{J.S.Dowker {\it Robin conditions on the Euclidean ball} MUTP/95/7; hep-th\break/9506042.} \reference{Dow9}{J.S.Dowker {\it Oddball determinants} MUTP/95/12; hep-th/9507096.} \reference{Dow10}{J.S.Dowker \pr{D28}{83}{3013}.} \reference{BandG}{P.B.Gilkey and T.P.Branson \tams{344}{94}{479}.} \reference{Schofield}{J.P.Schofield Ph.D.thesis, University of Manchester, (1991).} \reference{Barnesa}{E.W.Barnes {\it Trans. Camb. Phil. Soc.} {\bf 19} (1903) 374.} \reference{BandG2}{T.P.Branson and P.B.Gilkey {\it Comm. Partial Diff. Equations} {\bf 15} (1990) 245.} \reference{Pathria}{R.K.Pathria {\it Suppl.Nuovo Cim.} {\bf 4} (1966) 276.} \reference{Baltes}{H.P.Baltes \prA{6}{72}{2252}.} \reference{Spivak}{M.Spivak {\it Differential Geometry} vols III, IV, Publish or Perish, Boston, 1975.} \reference{Eisenhart}{L.P.Eisenhart {\it Differential Geometry}, Princeton University Press, Princeton, 1926.} \reference{Moss}{I.G.Moss \cqg{6}{89}{659}.} \reference{Barv}{A.O.Barvinsky, Yu.A.Kamenshchik and I.P.Karmazin \aop {219} {92}{201}.} \reference{Kam}{Yu.A.Kamenshchik and I.V.Mishakov \prD{47}{93}{1380}.} \reference{KandM}{Yu.A.Kamenshchik and I.V.Mishakov {\it Int. J. Mod. Phys.} {\bf A7} (1992) 3265.} \reference{DandE}{P.D.D'Eath and G.V.M.Esposito \prD{43}{91}{3234}.} \reference{DandE2}{P.D.D'Eath and G.V.M.Esposito \prD{44}{91}{1713}.} \reference{Rich}{K.Richardson \jfa{122}{94}{52}.} \reference{Osgood}{B.Osgood, R.Phillips and P.Sarnak \jfa{80}{88}{148}.} \reference{BCY}{T.P.Branson, S.-Y. A.Chang and P.C.Yang \cmp{149}{92}{241}.} \reference{Vass}{D.V.Vassilevich.{\it Vector fields on a disk with mixed boundary conditions} gr-qc /9404052.} \reference{MandP}{I.Moss and S.Poletti \pl{B333}{94}{326}.} \reference{Kam2}{G.Esposito, A.Y.Kamenshchik, I.V.Mishakov and G.Pollifrone \prD{50}{94}{6329}.} \reference{Aurell1}{E.Aurell and P.Salomonson \cmp{165}{94}{233}.} \reference{Aurell2}{E.Aurell and P.Salomonson {\it Further results on functional determinants of laplacians on simplicial complexes} hep-th/9405140.} \reference{BandO}{T.P.Branson and B.\O rsted \pams{113}{91}{669}.} \reference{Elizalde1}{E.Elizalde, \jmp{35}{94}{3308}.} \reference{BandK}{M.Bordag and K.Kirsten {\it Heat-kernel coefficients of the Laplace operator on the 3-dimensional ball} hep-th/9501064.} \reference{Waechter}{R.T.Waechter \pcps{72}{72}{439}.} \reference{GRV}{S.Guraswamy, S.G.Rajeev and P.Vitale {\it O(N) sigma-model as a three dimensional conformal field theory}, Rochester preprint UR-1357.} \reference{CandC}{A.Capelli and A.Costa \np {314}{89}{707}.} \reference{IandZ}{C.Itzykson and J.-B.Zuber \np{275}{86}{580}.} \reference{BandH}{M.V.Berry and C.J.Howls \prs {447}{94}{527}.} \reference{DandW}{A.Dettki and A.Wipf \np{377}{92}{252}.} \reference{Weisbergerb} {W.I.Weisberger \cmp{112}{87}{633}.} \reference{Voros}{A.Voros \cmp{110}{87}{110}.} \reference{Pockels}{F.Pockels {\it \"Uber die partielle Differentialgleichung $\Delta u+k^2u=0$}, B.G.Teubner, Leipzig 1891.} \reference{Kober}{H.Kober \mz{39}{1935}{609}.} \reference{Watson2}{G.N.Watson \qjm{2}{31}{300}.} \reference{DandC1}{J.S.Dowker and R.Critchley \prD {13}{76}{3224}.} \reference{DandC2}{J.S.Dowker and R.Critchley \prD {13}{76}{224}.} \reference{Lamb}{H.Lamb \pm{15}{1884}{270}.} \reference{EandR}{E.Elizalde and A.Romeo International J. of Math. and Phys. {\bf13} (1994) 453} \reference{DandA}{J.S.Dowker and J.S.Apps \cqg{12}{95}{1363}.} \reference{DandA2}{J.S.Dowker and J.S.Apps, {\it Functional determinants on certain domains}. To appear in the Proceedings of the 6th Moscow Quantum Gravity Seminar held in Moscow, June 1995; hep-th/9506204.} \reference{Watson1}{G.N.Watson {\it Theory of Bessel Functions} Cambridge University Press, Cambridge, 1944.} \reference{BGKE}{M.Bordag, B.Geyer, K.Kirsten and E.Elizalde, {\it Zeta function determinant of the Laplace operator on the D-dimensional ball} UB-ECM-PF 95/10; hep-th /9505157.} \reference{MandO}{W.Magnus and F.Oberhettinger {\it Formeln und S\"atze} Springer-Verlag, Berlin, 1948.} \reference{Olver}{F.W.J.Olver {\it Phil.Trans.Roy.Soc} {\bf A247} (1954) 328.} \reference{Hurt}{N.E.Hurt {\it Geometric Quantization in action} Reidel, Dordrecht, 1983.} \reference{Esposito}{G.Esposito {\it Quantum Gravity, Quantum Cosmology and Lorentzian Geometry}, Lecture Notes in Physics, Monographs, Vol. m12, Springer-Verlag, Berlin 1994.} \reference{Louko}{J.Louko \prD{38}{88}{478}.} \reference{Schleich} {K.Schleich \prD{32}{85}{1989}.} \reference{BEK}{M.Bordag, E.Elizalde and K.Kirsten {\it Heat kernel coefficients of the Laplace operator on the D-dimensional ball} UB-ECM-PF 95/3; hep-th/9503023.} \reference{ELZ}{E.Elizalde, S.Leseduarte and S.Zerbini.} \reference{BGV}{T.P.Branson, P.B.Gilkey and D.V.Vassilevich {\it The Asymptotics of the Laplacian on a manifold with boundary} II, hep-th/9504029.} \reference{Erdelyi}{A.Erdelyi,W.Magnus,F.Oberhettinger and F.G.Tricomi {\it Higher Transcendental Functions} Vol.I McGraw-Hill, New York, 1953.} \reference{Quine}{J.R.Quine, S.H.Heydari and R.Y.Song \tams{338}{93}{213}.} \reference{Dikii}{L.A.Dikii {\it Usp. Mat. Nauk.} {\bf13} (1958) 111.} \reference{DandH}{P.D.D'Eath and J.J.Halliwell \prd{35}{87}{1100}.} \reference{KandC}{K.Kirsten and G.Cognola, {\it Heat-kernel coefficients and functional determinants for higher spin fields on the ball} UTF354. Aug. 1995.} \reference{Louko}{J.Louko \prD{38}{88}{478}.} \reference{MandP}{I.G.Moss and S.J.Poletti \pl{B333}{94}{326}.} \reference{MandP2}{I.G.Moss and S.J.Poletti \np{341}{90}{155}.} \reference{Luck}{H.C.Luckock \jmp{32}{91}{1755}.} \reference{Poletti}{S.J.Poletti \pl{B249}{90}{355}.} \reference{gilk}{P.B.Gilkey {\it Invariant theory, the heat equation and the Atiyah-Singer index theorem}, Publish or Perish, Wilmington, DE, 1984.} \reference{deB}{Louis de Broglie {\it Probl\`emes de propagation guide\'es des ondes electromagnetiques} 2me. \'Ed. Gauthier-Villars, Paris, 1951.} \end{putreferences} \bye
\section*{Figure captions} \begin{figure}[h] \caption{The anisotropy coefficient $B_\eta$ for the $\eta$--channel at initial energies from 1.26 to 4.9~GeV.} \label{fig1} \caption{ The cross sections $d\sigma/dM$ for $pp$ and $pn$ interactions at 1.26 and 2.1~GeV bombarding energies. The ``$\eta$'' denotes the contribution of the $\eta$--channel, the ``$\Delta$'' labels the contribution of the $\Delta$--resonance term, while ``$Br$'' denotes the bremsstrahlung channel. The dashed curves are the sum of bremsstrahlung and $\Delta$--channel with interference.} \label{fig2} \caption{The weighted anisotropy coefficients $<B_i(M)>$ for $pp$ and $pn$ collisions at 1.26 and 2.1~GeV. The notation is the same as in Fig.~\protect\ref{fig2}. The solid line denoted by ``$all$'' represents the sum of bremsstrahlung, $\Delta$--channel with interference and $\eta$--decay contributions. } \label{fig3} \caption{The weighted anisotropy coefficients $<B_i(M)>$ for $pd$ interactions at 1.26 and 2.1~GeV. The notation is the same as in Fig.~\protect\ref{fig3}. } \label{fig4} \caption{The ratio of anisotropy coefficients for $pd$ to $pp$ reactions at energies of 1.26 and 2.1~GeV. The solid lines correspond to calculations taking into account the $\eta$--decay contribution while the dashed lines do not include the $\eta$--Dalitz decay.} \label{fig5} \end{figure} \end{document}
\section*{Figure Captions\markboth {FIGURECAPTIONS}{FIGURECAPTIONS}}\list {Figure \arabic{enumi}:\hfill}{\settowidth\labelwidth{Figure 999:} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endfigcap\endlist \relax \def\tablecap{\section*{Table Captions\markboth {TABLECAPTIONS}{TABLECAPTIONS}}\list {Table \arabic{enumi}:\hfill}{\settowidth\labelwidth{Table 999:} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endtablecap\endlist \relax \def\reflist{\section*{References\markboth {REFLIST}{REFLIST}}\list {[\arabic{enumi}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endreflist\endlist \relax \makeatletter \newcounter{pubctr} \def\@ifnextchar[{\@publist}{\@@publist}{\@ifnextchar[{\@publist}{\@@publist}} \def\@publist[#1]{\list {[\arabic{pubctr}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep \@nmbrlisttrue\def\@listctr{pubctr} \setcounter{pubctr}{#1}\addtocounter{pubctr}{-1}}} \def\@@publist{\list {[\arabic{pubctr}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep \@nmbrlisttrue\def\@listctr{pubctr}}} \let\endpublist\endlist \relax \makeatother \def\hskip -.1cm \cdot \hskip -.1cm{\hskip -.1cm \cdot \hskip -.1cm} \def\not\!{\not\!} \catcode`\@=11 \def\section{\@startsection {section}{1}{0pt}{-3.5ex plus -1ex minus -.2ex}{2.3ex plus .2ex}{\raggedright\large\bf}} \catcode`\@=12 \def\mbf#1{\hbox{\boldmath $#1$}} \newskip\humongous \humongous=0pt plus 1000pt minus 1000pt \def\mathsurround=0pt{\mathsurround=0pt} \def\eqalign#1{\,\vcenter{\openup1\jot \mathsurround=0pt \ialign{\strut \hfil$\displaystyle{##}$&$ \displaystyle{{}##}$\hfil\crcr#1\crcr}}\,} \newif\ifdtup \def\panorama{\global\dtuptrue \openup1\jot \mathsurround=0pt \everycr{\noalign{\ifdtup \global\dtupfalse \vskip-\lineskiplimit \vskip\normallineskiplimit \else \penalty\interdisplaylinepenalty \fi}}} \def\eqalignno#1{\panorama \tabskip=\humongous \halign to\displaywidth{\hfil$\displaystyle{##}$ \tabskip=0pt&$\displaystyle{{}##}$\hfil \tabskip=\humongous&\llap{$##$}\tabskip=0pt \crcr#1\crcr}} \def\hangindent3\parindent{\hangindent3\parindent} \def\oldreffmt#1{\rlap{[#1]} \hbox to 2\parindent{}} \def\oldref#1{\par\noindent\hangindent3\parindent \oldreffmt{#1} \ignorespaces} \def\hangindent=1.25in{\hangindent=1.25in} \def\figfmt#1{\rlap{Figure {#1}} \hbox to 1in{}} \def\fig#1{\par\noindent\hangindent=1.25in \figfmt{#1} \ignorespaces} \def\hbox{\it i.e.}} \def\etc{\hbox{\it etc.}{\hbox{\it i.e.}} \def\etc{\hbox{\it etc.}} \def\hbox{\it e.g.}} \def\cf{\hbox{\it cf.}{\hbox{\it e.g.}} \def\cf{\hbox{\it cf.}} \def\hbox{\it et al.}{\hbox{\it et al.}} \def\hbox{---}{\hbox{---}} \def\mathop{\rm cok}{\mathop{\rm cok}} \def\mathop{\rm tr}{\mathop{\rm tr}} \def\mathop{\rm Tr}{\mathop{\rm Tr}} \def\mathop{\rm Im}{\mathop{\rm Im}} \def\mathop{\rm Re}{\mathop{\rm Re}} \def\mathop{\bf R}{\mathop{\bf R}} \def\mathop{\bf C}{\mathop{\bf C}} \def\lie{\hbox{\it \$}} \def\partder#1#2{{\partial #1\over\partial #2}} \def\secder#1#2#3{{\partial^2 #1\over\partial #2 \partial #3}} \def\bra#1{\left\langle #1\right|} \def\ket#1{\left| #1\right\rangle} \def\VEV#1{\left\langle #1\right\rangle} \let\vev\VEV \def\gdot#1{\rlap{$#1$}/} \def\abs#1{\left| #1\right|} \def\pri#1{#1^\prime} \def\raisebox{-.4ex}{\rlap{$\sim$}} \raisebox{.4ex}{$<$}{\raisebox{-.4ex}{\rlap{$\sim$}} \raisebox{.4ex}{$<$}} \def\raisebox{-.4ex}{\rlap{$\sim$}} \raisebox{.4ex}{$>$}{\raisebox{-.4ex}{\rlap{$\sim$}} \raisebox{.4ex}{$>$}} \def\contract{\makebox[1.2em][c]{ \mbox{\rule{.6em}{.01truein}\rule{.01truein}{.6em}}}} \def{1\over 2}{{1\over 2}} \def\begin{equation}{\begin{equation}} \def\end{equation}{\end{equation}} \def\underline{\underline} \def\begin{eqnarray}{\begin{eqnarray}} \def\lrover#1{ \raisebox{1.3ex}{\rlap{$\leftrightarrow$}} \raisebox{ 0ex}{$#1$}} \def\com#1#2{ \left[#1, #2\right]} \def\end{eqnarray}{\end{eqnarray}} \def\:\raisebox{1.3ex}{\rlap{$\vert$}}\!\rightarrow{\:\raisebox{1.3ex}{\rlap{$\vert$}}\!\rightarrow} \def\longbent{\:\raisebox{3.5ex}{\rlap{$\vert$}}\raisebox{1.3ex}% {\rlap{$\vert$}}\!\rightarrow} \def\onedk#1#2{ \begin{equation} \begin{array}{l} #1 \\ \:\raisebox{1.3ex}{\rlap{$\vert$}}\!\rightarrow #2 \end{array} \end{equation} } \def\dk#1#2#3{ \begin{equation} \begin{array}{r c l} #1 & \rightarrow & #2 \\ & & \:\raisebox{1.3ex}{\rlap{$\vert$}}\!\rightarrow #3 \end{array} \end{equation} } \def\dkp#1#2#3#4{ \begin{equation} \begin{array}{r c l} #1 & \rightarrow & #2#3 \\ & & \phantom{\; #2}\:\raisebox{1.3ex}{\rlap{$\vert$}}\!\rightarrow #4 \end{array} \end{equation} } \def\bothdk#1#2#3#4#5{ \begin{equation} \begin{array}{r c l} #1 & \rightarrow & #2#3 \\ & & \:\raisebox{1.3ex}{\rlap{$\vert$}}\raisebox{-0.5ex}{$\vert$}% \phantom{#2}\!\:\raisebox{1.3ex}{\rlap{$\vert$}}\!\rightarrow #4 \\ & & \:\raisebox{1.3ex}{\rlap{$\vert$}}\!\rightarrow #5 \end{array} \end{equation} } \hyphenation{anom-a-ly} \hyphenation{comp-act-ifica-tion} \def\ap#1#2#3{ {\it Ann. Phys. (NY) }{\bf #1}, #2 (19#3)} \def\apj#1#2#3{ {\it Astrophys. J. }{\bf #1}, #2 (19#3)} \def\apjl#1#2#3{ {\it Astrophys. J. Lett. }{\bf #1}, #2 (19#3)} \def\app#1#2#3{ {\it Acta Phys. Polon. }{\bf #1}, #2 (19#3)} \def\ar#1#2#3{ {\it Ann. Rev. Nucl. and Part. Sci. }{\bf #1}, #2 (19#3)} \def\com#1#2#3{ {\it Comm. Math. Phys. }{\bf #1}, #2 (19#3)} \def\ib#1#2#3{ {\it ibid. }{\bf #1}, #2 (19#3)} \def\nat#1#2#3{ {\it Nature (London) }{\bf #1}, #2 (19#3)} \def\nc#1#2#3{ {\it Nuovo Cim. }{\bf #1}, #2 (19#3)} \def\np#1#2#3{ {\it Nucl. Phys. }{\bf #1}, #2 (19#3)} \def\pl#1#2#3{ {\it Phys. Lett. }{\bf #1}, #2 (19#3)} \def\pr#1#2#3{ {\it Phys. Rev. }{\bf #1}, #2 (19#3)} \def\prep#1#2#3{ {\it Phys. Rep. }{\bf #1}, #2 (19#3)} \def\prl#1#2#3{ {\it Phys. Rev. Lett. }{\bf #1}, #2 (19#3)} \def\pro#1#2#3{ {\it Prog. Theor. Phys. }{\bf #1}, #2 (19#3)} \def\rmp#1#2#3{ {\it Rev. Mod. Phys. }{\bf #1}, #2 (19#3)} \def\sp#1#2#3{ {\it Sov. Phys.-Usp. }{\bf #1}, #2 (19#3)} \def\sjn#1#2#3{ {\it Sov. J. Nucl. Phys. }{#1}, #2 (19#3)} \def\srv#1#2#3{ {\it Surv. High Energy Phys. }{#1}, #2 (19#3)} \defthese proceedings{these proceedings} \def\zp#1#2#3{ {\it Zeit. fur Physik }{\bf #1}, #2 (19#3)} \catcode`\@=11 \def\eqnarray{\stepcounter{equation}\let\@currentlabel=\thesection.\arabic{equation}} \global\@eqnswtrue \global\@eqcnt\z@\tabskip\@centering\let\\=\@eqncr \gdef\@@fix{}\def\eqno##1{\gdef\@@fix{##1}}% $$\halign to \displaywidth\bgroup\@eqnsel\hskip\@centering $\displaystyle\tabskip\z@{##}$&\global\@eqcnt\@ne \hskip 2\arraycolsep \hfil${##}$\hfil &\global\@eqcnt\tw@ \hskip 2\arraycolsep $\displaystyle\tabskip\z@{##}$\hfil \tabskip\@centering&\llap{##}\tabskip\z@\cr} \def\@@eqncr{\let\@tempa\relax \ifcase\@eqcnt \def\@tempa{& & &}\or \def\@tempa{& &} \else \def\@tempa{&}\fi \@tempa \if@eqnsw\@eqnnum\stepcounter{equation}\else\@@fix\gdef\@@fix{}\fi \global\@eqnswtrue\global\@eqcnt\z@\cr} \catcode`\@=12 \font\tenbifull=cmmib10 \font\tenbimed=cmmib10 scaled 800 \font\tenbismall=cmmib10 scaled 666 \textfont9=\tenbifull \scriptfont9=\tenbimed \scriptscriptfont9=\tenbismall \def\fam9 {\fam9 } \def\fam=9{\mathchar"710B } {\fam=9{\mathchar"710B } } \def\fam=9{\mathchar"710C } {\fam=9{\mathchar"710C } } \def\fam=9{\mathchar"710D } {\fam=9{\mathchar"710D } } \def\fam=9{\mathchar"710E } {\fam=9{\mathchar"710E } } \def\fam=9{\mathchar"710F } {\fam=9{\mathchar"710F } } \def\fam=9{\mathchar"7110 } {\fam=9{\mathchar"7110 } } \def\fam=9{\mathchar"7111 } {\fam=9{\mathchar"7111 } } \def\fam=9{\mathchar"7112 } {\fam=9{\mathchar"7112 } } \def\fam=9{\mathchar"7113 } {\fam=9{\mathchar"7113 } } \def\fam=9{\mathchar"7114 } {\fam=9{\mathchar"7114 } } \def\fam=9{\mathchar"7115 } {\fam=9{\mathchar"7115 } } \def\fam=9{\mathchar"7116 } {\fam=9{\mathchar"7116 } } \def\fam=9{\mathchar"7117 } {\fam=9{\mathchar"7117 } } \def\fam=9{\mathchar"716F } {\fam=9{\mathchar"716F } } \def\fam=9{\mathchar"7118 } {\fam=9{\mathchar"7118 } } \def\fam=9{\mathchar"7119 } {\fam=9{\mathchar"7119 } } \def\fam=9{\mathchar"711A } {\fam=9{\mathchar"711A } } \def\fam=9{\mathchar"711B } {\fam=9{\mathchar"711B } } \def\fam=9{\mathchar"711C } {\fam=9{\mathchar"711C } } \def\fam=9{\mathchar"711D } {\fam=9{\mathchar"711D } } \def\fam=9{\mathchar"711E } {\fam=9{\mathchar"711E } } \def\fam=9{\mathchar"711F } {\fam=9{\mathchar"711F } } \def\fam=9{\mathchar"7120 } {\fam=9{\mathchar"7120 } } \def\fam=9{\mathchar"7121 } {\fam=9{\mathchar"7121 } } \def\fam=9{\mathchar"7122 } {\fam=9{\mathchar"7122 } } \def\fam=9{\mathchar"7123 } {\fam=9{\mathchar"7123 } } \def\fam=9{\mathchar"7124 } {\fam=9{\mathchar"7124 } } \def\fam=9{\mathchar"7125 } {\fam=9{\mathchar"7125 } } \def\fam=9{\mathchar"7126 } {\fam=9{\mathchar"7126 } } \def\fam=9{\mathchar"7127 } {\fam=9{\mathchar"7127 } } \def\fam=6{\mathchar"7000 } {\fam=6{\mathchar"7000 } } \def\fam=6{\mathchar"7001 } {\fam=6{\mathchar"7001 } } \def\fam=6{\mathchar"7002 } {\fam=6{\mathchar"7002 } } \def\fam=6{\mathchar"7003 } {\fam=6{\mathchar"7003 } } \def\fam=6{\mathchar"7004 } {\fam=6{\mathchar"7004 } } \def\fam=6{\mathchar"7005 } {\fam=6{\mathchar"7005 } } \def\fam=6{\mathchar"7006 } {\fam=6{\mathchar"7006 } } \def\fam=6{\mathchar"7007 } {\fam=6{\mathchar"7007 } } \def\fam=6{\mathchar"7008 } {\fam=6{\mathchar"7008 } } \def\fam=6{\mathchar"7009 } {\fam=6{\mathchar"7009 } } \def\fam=6{\mathchar"700A } {\fam=6{\mathchar"700A } } \def\fam=9{\mathchar"700A } {\fam=9{\mathchar"700A } } \def\fam=9{\mathchar"7000 } {\fam=9{\mathchar"7000 } } \def\fam=9{\mathchar"7001 } {\fam=9{\mathchar"7001 } } \def\fam=9{\mathchar"7002 } {\fam=9{\mathchar"7002 } } \def\fam=9{\mathchar"7003 } {\fam=9{\mathchar"7003 } } \def\fam=9{\mathchar"7004 } {\fam=9{\mathchar"7004 } } \def\fam=9{\mathchar"7005 } {\fam=9{\mathchar"7005 } } \def\fam=9{\mathchar"7006 } {\fam=9{\mathchar"7006 } } \def\fam=9{\mathchar"7007 } {\fam=9{\mathchar"7007 } } \def\fam=9{\mathchar"7008 } {\fam=9{\mathchar"7008 } } \def\fam=9{\mathchar"7009 } {\fam=9{\mathchar"7009 } } \def\fam=9{\mathchar"700A } {\fam=9{\mathchar"700A } } \relax \def\double{ \renewcommand{1.2}{2} \large \normalsize } \def\single { \renewcommand{1.2}{1} \large \normalsize } \def1.2{1.2} \def\ thefootnote{\fnsymbol{footnote}} \def\widetilde{\widetilde} \def\widehat{\widehat} \renewcommand{\fam=9{\mathchar"7115 } }{\mbox{\boldmath $\lambda$}} \renewcommand{\fam=9{\mathchar"7110 } }{\mbox{\boldmath $\zeta$}} \begin{document} \begin{titlepage} \begin{center} \today \hfill LBL-37343\\ \hfill UCB-PTH-95/17\\ \hfill hep-ph/9508288 \\ \vskip .1in \vskip .25in {\large \bf Flavor Mixing Signals For Realistic\\ Supersymmetric Unification.} \footnote{This work was supported in part by the Director, Office of Energy Research, Office of High Energy and Nuclear Physics, Division of High Energy Physics of the U.S. Department of Energy under Contract DE-AC03-76SF00098 and in part by the National Science Foundation under grant PHY-90-21139.} \vskip .25in {\bf Nima Arkani-Hamed\\ Hsin-Chia Cheng}\\ and\\ {\bf L.J. Hall}\\ \vskip .20in {\em Theoretical Physics Group\\ Lawrence Berkeley Laboratory\\ and\\ Department of Physics\\ University of California\\ Berkeley, California 94720} \end{center} \vskip .25in \begin{abstract} The gauge interactions of any supersymmetric extension of the standard model involve new flavor mixing matrices. The assumptions involved in the construction of minimal supersymmetric models, both $SU(3) \times SU(2) \times U(1)$ and grand unified theories, force a large degree of triviality on these matrices. However, the requirement of realistic quark and lepton masses in supersymmetric grand unified theories forces these matrices to be non-trivial. This leads to important new dominant contributions to the neutron electric dipole moment and to the decay mode $p \to K^o\mu^+$, and suggests that there may be important weak scale radiative corrections to the Yukawa coupling matrix of the up quarks. The lepton flavor violating signal $\mu \to e\gamma$ is studied in these theories when $\tan\beta$ is sufficiently large that radiative effects of couplings other than $\lambda_t$ must be included. The naive expectation that large $\tan\beta$ will force sleptons to unacceptably large masses is not borne out: radiative suppressions to the leptonic flavor mixing angles allow regions where the sleptons are as light as 300 GeV, provided the top Yukawa coupling in the unified theory is near the minimal value consistent with $m_t$. \end{abstract} \end{titlepage} \renewcommand{\thepage}{\roman{page}} \setcounter{page}{2} \mbox{ } \vskip 1in \begin{center} {\bf Disclaimer} \end{center} \vskip .2in \begin{scriptsize} \begin{quotation} This document was prepared as an account of work sponsored by the United States Government. While this document is believed to contain correct information, neither the United States Government nor any agency thereof, nor The Regents of the University of California, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial products process, or service by its trade name, trademark, manufacturer, or otherwise, does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof, or The Regents of the University of California. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof of The Regents of the University of California and shall not be used for advertising or product endorsement purposes. \end{quotation} \end{scriptsize} \vskip 2in \begin{center} \begin{small} {\it Lawrence Berkeley Laboratory is an equal opportunity employer.} \end{small} \end{center} \newpage \renewcommand{\thepage}{\arabic{page}} \setcounter{page}{1} \noindent{\bf I. Introduction} \medskip It has recently been demonstrated that flavor and CP violation provide an important new probe of supersymmetric grand unified theories [1-4]. These new signals, such as $\mu\to e\gamma$ and the electron electric dipole moment $d_e$, are complementary to the classic tests of proton decay, neutrino masses and quark and charged lepton mass relations. The classic tests are very dependent on the flavor interactions and symmetry breaking sector of the unified model: it is only too easy to construct models in which these signals are absent or unobservable. However, they are insensitive to the hardness scale, $\Lambda_H$, of supersymmetry breaking.\footnote{This is the highest scale at which supersymmetry breaking squark and gluino masses appear in the theory as local interactions.} On the other hand, the new flavor and CP violating signals are relatively insensitive to the form of the flavor interactions and unified gauge symmetry breaking, but are absent if the hardness scale, $\Lambda_H$, falls beneath the unified scale, $M_G$. The signals are generated by the unified flavor interactions leaving an imprint on the form of the soft supersymmetry breaking operators \cite{HKR}, which is only possible if supersymmetry breaking is present in the unified theory at scales above $M_G$. The flavor and CP violating signals have been computed in the minimal $SU(5)$ and $SO(10)$ models for leptonic [1-3] and hadronic processes \cite{BHS2}, for moderate values of tan$\beta$, the ratio of the two Higgs vacuum expectation values. While rare muon decays provide an important probe of $SU(5)$, it is the $SO(10)$ theory which is most powerfully tested. If the hardness scale for supersymmetry breaking is large enough, as in the popular supergravity models, it may be possible for the minimal $SO(10)$ theory to be probed throughout the interesting range of superpartner masses by searches for $\mu \to e\gamma$ and $d_e$. The flavor changing and CP violating probes of $SO(10)$ are sufficiently powerful to warrant an exploration of consequences for non-minimal models, which is the subject of this paper. In particular, we study $SO(10)$ theories in which (I) {\it{The Yukawa interactions are non-minimal.}} In the minimal model the quarks and leptons lie in three 16's and the two Higgs doublets $H_U$ and $H_D$ lie in two 10 dimensional representations $10_U$ and $10_D$. The quark and charged lepton masses are assumed to arise from the interactions $16 {\fam=9{\mathchar"7115 } }_U 16\ 10_U + 16 {\fam=9{\mathchar"7115 } }_D 16\ 10_D$. This model is a useful fiction: it is very simple to work with, but leads to the mass relation $m_e/m_\mu = m_d/m_s$, which is in error by an order of magnitude. It is clearly necessary to introduce a mechanism to insert $SO(10)$ breaking into the Yukawa interactions. The simplest way to achieve this is to assume that at the unification scale, $M_G$, some of the Yukawa interactions arise from higher dimensional operators involving fields $A$ which break the $SO(10)$ symmetry group. This implies that ${\fam=9{\mathchar"7115 } }_{U,D} \to {\fam=9{\mathchar"7115 } }_{U,D}(A)$. Every realistic model of $SO(10)$ which has been constructed has this form; hence one should view this generalization of the minimal model as a necessity. (II) {\it{The ratio of electroweak VEV's,}} $\tan\beta = v_U/v_D$, {\it{ is allowed to be large,}} $\approx m_t/m_b$. This is certainly not a necessity; to the contrary, a simple extrapolation of the results of \cite{BHS} to such large values of $\tan\beta$ suggests that it is already excluded by the present limit on $\mu \to e\gamma$. The case of large $\tan\beta$ in $SO(10)$ has received much attention \cite{ALS,HRS,RS,ADHRS} partly because it has important ramifications for the origin of $m_t/m_b = (\lambda_t/\lambda_b) \tan\beta$. To what extent is this puzzling large ratio to be understood as a large hierarchy of Yukawa couplings, and to what extent in terms of a large value for $\tan\beta$? If the third generation masses arise from a single interaction of the form $16_3 16_3 10$ it is possible to predict $m_t$ using $m_b$ and $m_\tau$ as input \cite{ALS}, providing the theory is perturbative up to $M_G$. The prediction is $175 \pm 10 $ GeV \cite{HRS}, and requires $\tan\beta \approx m_t/m_b$. In this paper we investigate whether this intriguing possibility is excluded by the $\mu \to e \gamma$ signal; or, more correctly, we determine whether it requires a soft origin for supersymmmetry breaking, making it incompatible with the standard supergravity scenario \cite{CAN}. In the next section we show that $SO(10)$ models with ${\fam=9{\mathchar"7115 } } \to {\fam=9{\mathchar"7115 } } (A)$ possess new gaugino mixing matrices in the up-quark sector, which did not arise in the minimal models. In section III we set our notation for the supersymmetric standard model with arbitrary gaugino mixing matrices, and we show which mixing matrices are expected from unified models according to the gauge group and the value of $\tan\beta$. In section IV we describe the new phenomenological signatures which are generated by the gaugino mixing matrices in the up sector; these signatures are generic to all models with Yukawa interactions generated from higher dimensional operators. The consequences of large $\tan\beta$ for the flavor and CP violating signatures are analyzed analytically in section V and numerically in section VI. The analysis of the first five sections applies to a wide class of models. In section VII we illustrate the results in the particular models introduced by Anderson et al \cite{ADHRS}. As well as providing illustrations, these models have features unique to themselves. Conclusions are drawn in section VIII. \newpage \noindent{\bf II. New Flavor Mixing in the Up Sector} \medskip In [1-4] flavor and CP violating signals are studied in minimal $SU(5)$ and $SO(10)$ models with moderate $\tan\beta$. In these models the radiative corrections to the scalar mass matrices are dominated by the top quark Yukawa coupling $\lambda_t$ of the unified theory, so the scalar mass matrices tend to align with the up-type Yukawa coupling matrix and all non-trivial flavor mixing matrices are simply related to the KM matrix. However, as mentioned above, the minimal models do not give realistic fermion masses. One has to insert $SO(10)$ breaking into the Yukawa interactions. The simplest way to achieve this is to assume that the light fermion masses come from the non-renormalizable operators $$ \lambda'_{ij} 16_i{A_1\over M_1}{A_2\over M_2}\ldots {A_\ell\over M_\ell} 10 {A_{\ell+1}\over M_{\ell +1}} \ldots {A_n\over M_{n}} 16_j,\eqno(2.1) $$ where the $16_i$'s contain the three low energy families, 10 contains the Higgs doublets, and $A$'s are adjoint fields with vacuum expectation values (VEV's) which break the $SO(10)$ gauge group. After substituting in the VEV's of the adjoints, they become the usual Yukawa interactions with different Clebsch factors associated with Yukawa couplings of fields with different quantum numbers. For example in the models introduced by Anderson et al. \cite{ADHRS}, (hereafter referred to as ADHRS models) $$ {\fam=9{\mathchar"7115 } }_U = \pmatrix{0&z_uC &0\cr z'_uC & y_uE &x_uB\cr 0 & x'_uB &A}, {\fam=9{\mathchar"7115 } }_D = \pmatrix{ 0 & z_dC & C\cr z'_d & y_dE & x_dB\cr 0 & x'_dB &A}, {\fam=9{\mathchar"7115 } }_E = \pmatrix{ 0 & z_eC & 0\cr z'_eC & y_eE & x_eB\cr 0 & x'_eB & A},\eqno(2.2) $$ where the $x,y,z$'s are Clebsch factors arising from the VEV's of the adjoint fields. Thus realistic fermion masses and mixings can be obtained. The radiative corrections to the soft SUSY-breaking operators above $M_G$ are now more complicated. {}From the interactions (2.1) the following soft supersymmetry breaking operators are generated: $$ \lambda^\dagger_{ik}(A) m^2_{k\ell}(A) \lambda_{\ell j}(A) \; \phi^\dagger_i\phi_j,\eqno(2.3) $$ where $\phi_i, \phi_j$ are scalar components of the superfields, and $\lambda_{ij}(A)$ are adjoint dependent couplings, $\lambda(A) = \lambda' {A_1\over M_1} \ldots {A_n\over M_n}$. After the adjoints take their VEV's, the $m^2_{k\ell} (A)$ become the usual soft scalar masses. If we ignore the wavefunction renormalization of the adjoint fields (which is valid in the one-loop approximation), this is the same as if we had replaced the adjoints by their VEV's all the way up to the ultraheavy scale where the ultraheavy fields are integrated out, and treated these nonrenormalizable operators as the usual Yukawa interactions and scalar mass operators. This is a convenient way of thinking and we will use it in the rest of the paper. Above the GUT scale, in addition to the Yukawa interactions which give the fermion masses $$ Q{\fam=9{\mathchar"7115 } }_U U^cH_U, \; Q{\fam=9{\mathchar"7115 } }_D D^c H_D, \; E^c {\fam=9{\mathchar"7115 } }_E L H_D, \eqno(2.4) $$ the operators (2.1) also lead to $$ \eqalignno{ Q{\fam=9{\mathchar"7115 } }_{qq} Q H_{U_3}, &E^c {\fam=9{\mathchar"7115 } }_{eu}U^c H_{U_3}, \; N{\fam=9{\mathchar"7115 } }_{nd}D^cH_{U_3},\cr Q \lambda_{q\ell} L H_{D_3}, &U^c{\fam=9{\mathchar"7115 } }_{ud}D^c H_{D_3}, N{\fam=9{\mathchar"7115 } }_{n\ell} LH_U,&(2.5)\cr} $$ where $H_{U_3}, H_{D_3}$ are the triplet partners of the two Higgs doublets $H_U$ and $H_D$. Each Yukawa matrix has different Clebsch factors associated with its elements, so they can not be diagonalized in the same basis. The scalar mass matrices receive radiative corrections from Yukawa interactions of both (2.4) and (2.5), which, in the one-loop approximation, take the form $$ \eqalignno{ \Delta{\bf{m}}^2_Q &\propto {\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U + {\fam=9{\mathchar"7115 } }_D{\fam=9{\mathchar"7115 } }_D^\dagger + 2{\fam=9{\mathchar"7115 } }_{qq} {\fam=9{\mathchar"7115 } }^\dagger_{qq}+ {\fam=9{\mathchar"7115 } }_{q\ell} {\fam=9{\mathchar"7115 } }^\dagger_{q\ell},\cr \Delta{\bf{m}}^2_U &\propto 2{\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_U + {\fam=9{\mathchar"7115 } }^\dagger_{eu}{\fam=9{\mathchar"7115 } }_{eu} + 2 {\fam=9{\mathchar"7115 } }_{ud} {\fam=9{\mathchar"7115 } }^\dagger_{ud},\cr \Delta{\bf{m}}^2_D &\propto 2{\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_D + {\fam=9{\mathchar"7115 } }^\dagger_{nd}{\fam=9{\mathchar"7115 } }_{nd} + 2{\fam=9{\mathchar"7115 } }_{ud}^\dagger {\fam=9{\mathchar"7115 } }_{ud},\cr \Delta{\bf{m}}^2_L &\propto {\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E + 3{\fam=9{\mathchar"7115 } }^\dagger_{q\ell} {\fam=9{\mathchar"7115 } }_{q\ell} + {\fam=9{\mathchar"7115 } }^\dagger_{n\ell}{\fam=9{\mathchar"7115 } }_{n\ell},\cr \Delta{\bf{m}}^2_E &\propto 2{\fam=9{\mathchar"7115 } }_E{\fam=9{\mathchar"7115 } }^\dagger_E + 3{\fam=9{\mathchar"7115 } }_{eu}{\fam=9{\mathchar"7115 } }^\dagger_{eu}.&(2.6) \cr} $$ In the minimal $SO(10)$ model, scalar mass renormalizations above $M_G$ arise from a single matrix ${\fam=9{\mathchar"7115 } }_U$. It is therefore possible to choose a ``U-basis'' in which the scalings are purely diagonal. This is clearly not possible in the general models. All scalar mass matrices and Yukawa matrices are in general diagonalized in different bases. Therefore, flavor mixing matrices should appear in all gaugino vertices, including in the up-quark sector (where they are trivial in the minimal models studied in [1-4]). The up-type quark-squark-gaugino flavor mixing is a novel feature of the general models. Its consequences will be discussed in Sec. IV. Also, the flavor mixing matrices are no longer simply the KM matrix. They are model dependent and are different for different types of quarks and charged leptons, and are fully described in the next section. \newpage \noindent {\bf III. Flavor Mixing Matrices in General Superymmetric Standard Models.} \medskip In this section we set our notation for the gaugino flavor mixing matrices in the supersymmetric theory below $M_G$, taken to have minimal field content. We also give general expectations for these matrices in a wide variety of unified theories. The most general scalar masses are $6 \times 6$ matrices for squarks and charged sleptons and $3 \times 3$ matrix for sneutrinos, $$ \eqalignno{ {\bf{m}}^2_{{U}} &= \pmatrix{{\bf{ m}}^2_{{U}_L} & ({\fam=9{\mathchar"7110 } }_U+{\fam=9{\mathchar"7115 } }_U\mu\cot \beta)v_U\cr ({\fam=9{\mathchar"7110 } }^\dagger_U + {\fam=9{\mathchar"7115 } }^\dagger_U\mu\cot\beta)v_U & {\bf{m}}^2_{{U}_R} },\cr {\bf{m}}^2_{{D}} &= \pmatrix{ {\bf{m}}^2_{{D}_L} & ({\fam=9{\mathchar"7110 } }_D +{\fam=9{\mathchar"7115 } }_D\mu\tan\beta)v_D\cr ({\fam=9{\mathchar"7110 } }^\dagger_D + {\fam=9{\mathchar"7115 } }^\dagger_D \mu\tan\beta) v_D & {\bf{m}}^2_{{D}_R} },\cr {\bf{m}}^2_{{E}} &= \pmatrix{ {\bf{m}}^2_{{E}_L} & ({\fam=9{\mathchar"7110 } }_E + {\fam=9{\mathchar"7115 } }_E\mu\tan\beta)v_D\cr ({\fam=9{\mathchar"7110 } }^\dagger_E + {\fam=9{\mathchar"7115 } }^\dagger_E\mu\tan\beta)v_D & {\bf{m}}^2_{{E}_R} }, \cr {\bf{m}}^2_{{\nu}} &= \left( {\bf{m}}^2_{{\nu}_{ij}} \right),&(3.1)\cr} $$ where ${\bf{m}}^2_{{U}_L}, {\bf{m}}^2_{{D}_L}, {\bf{m}}^2_{{U}_R}, {\bf{m}}^2_{{D}_R}, {\bf{m}}^2_{{E}_L}, {\bf{m}}^2_{{E}_R}$ are $3 \times 3$ soft SUSY-breaking mass matrices for the left-handed and right-handed squarks and sleptons, and ${\fam=9{\mathchar"7110 } }_U, {\fam=9{\mathchar"7110 } }_D, {\fam=9{\mathchar"7110 } }_E$ are the trilinear soft SUSY-breaking terms. To calculate flavor violating processes, such as $\mu \to e\gamma$, one can diagonalize the mass matrix ${\bf{m}}^2_{{E}}$ by the $6\times 6$ unitary rotation matrix $V_E$ and ${\bf{m}}^2_{{\nu}}$ by the 3 $\times $ 3 unitary rotation $V_\nu$, $$ {\bf{m}}^2_{{E}} = V_E \overline{\bf{m}}^2_{{E}} V^\dagger_E, \hskip .25in {\bf{m}}^2_{{\nu}} = V_\nu \overline{\bf{m}}^2_{{\nu}} V^\dagger_\nu,\eqno(3.2) $$ where ${\bf{\overline{m}}}^2_{{E}}, {\bf{\overline{m}}}^2_{{\nu}}$ are diagonal. The amplitude for $\mu \to e\gamma$ is given by the diagrams in Fig.\ 1, summing up all the internal scalar mass eigenstates. If the entries in the scalar mass matrices are arbitrary, they generally give unacceptably large rates for flavor violating processes. From the experimental limits one expects that the first two generation scalar masses should be approximately degenerate and the chirality-changing mass matrices ${\fam=9{\mathchar"7110 } }_A$ should be approximately proportional to the corresponding Yukawa coupling matrices ${\fam=9{\mathchar"7115 } }_A$. In this paper we treat the chirality-conserving mass matrices and chirality-changing mass matrices separately, i.e., the mass eigenstates are assumed to be purely left-handed or right-handed, and the chirality-changing mass terms are treated as a perturbation. This may not be a good approximation for the third generation where the Yukawa couplings are large, the correct treatment will be used in the numerical studies of Sec.\ VI. The superpotential contains $$ W \supset Q^T {{\fam=9{\mathchar"7115 } }}_U U^c H_U +Q^T {{\fam=9{\mathchar"7115 } }}_D D^c {H}_D + E^{cT}{{\fam=9{\mathchar"7115 } }}_E L {H}_D,\eqno(3.3) $$ where ${{\fam=9{\mathchar"7115 } }}_U, {\fam=9{\mathchar"7115 } }_D, {{\fam=9{\mathchar"7115 } }}_E$ are the Yukawa coupling matrices which are diagonalized by the left and right rotations, $$ \eqalignno{ {\fam=9{\mathchar"7115 } }_U &= V^*_{U_L}\overline{\fam=9{\mathchar"7115 } }_U V^\dagger_{U_R},\cr {\fam=9{\mathchar"7115 } }_D &= V^*_{D_L}\overline{\fam=9{\mathchar"7115 } }_D V^\dagger_{D_R},\cr {\fam=9{\mathchar"7115 } }_E &= V^*_{E_R}\overline{\fam=9{\mathchar"7115 } }_EV^\dagger_{E_L}.&(3.4)\cr} $$ The soft SUSY-breaking interactions contain $$ \eqalignno{ \widetilde{Q}^\dagger {\bf{m}}^{2*}_Q \widetilde{Q} &+ \widetilde{U}^{c\dagger} {\bf{m}}^2_U \widetilde{U}^c + \widetilde{D}^{c\dagger} {\bf{m}}^2_D \widetilde{D}^c + \widetilde{L}^\dagger {\bf{m}}^2_L \widetilde{L} + \widetilde{E}^{c\dagger} {\bf{m}}^{2*}_E \widetilde{E}^c\cr &+ \widetilde{Q}^T {{\fam=9{\mathchar"7110 } }}_U \widetilde{U}^c H_U + \widetilde{Q}^T {{\fam=9{\mathchar"7110 } }}_D \widetilde{D}^c {H}_D + \widetilde{E}^{cT}{{\fam=9{\mathchar"7110 } }}_E \widetilde{L}{H}_D.&(3.5) \cr} $$ Because the trilinear terms should be approximately proportional to the Yukawa couplings, we write $$ {{\fam=9{\mathchar"7110 } }} = {{\fam=9{\mathchar"7110 } }}_0 + \Delta{{\fam=9{\mathchar"7110 } }} = A {\fam=9{\mathchar"7115 } } + {{\Delta}}{\fam=9{\mathchar"7110 } }.\eqno(3.6) $$ The soft-breaking mass matrices are diagonalized by: $$ \eqalignno{ {\bf{m}}^{2*}_Q &= U_Q \overline{\bf{m}}^{2*}_Q U^\dagger_Q, \;\; {\bf{m}}^2_U = U_U \overline{\bf{m}}^2_U U^\dagger_U, \;\; {\bf{m}}^2_D = U_D\overline{\bf{m}}^2_D U^\dagger_D, \cr {\bf{m}}^2_L &= U_L\overline{\bf{m}}^2_L U^\dagger_L, \;\; {\bf{m}}^{2*}_E = U_E\overline{\bf{m}}^{2*}_E U^\dagger_E , &(3.7)\cr} $$ $$ {{\Delta{\fam=9{\mathchar"7110 } }}}_U = V^{'*}_{U_L}\Delta \overline{{\fam=9{\mathchar"7110 } }}_U V^{'\dagger}_{U_R},\; {{\Delta\fam=9{\mathchar"7110 } }}_D = V^{'*}_{D_L} {{\Delta}}\overline{{\fam=9{\mathchar"7110 } }}_D V^{'\dagger}_{D_R},\; {{\Delta\fam=9{\mathchar"7110 } }}_E = V^{'*}_{E_R} \Delta \overline{\fam=9{\mathchar"7110 } }_E V^{'\dagger}_{E_L}.\eqno(3.8) $$ In the mass eigenstate basis the rotation matrices $V, U$ appear in the gaugino couplings, \newpage $$ \eqalignno{ {\cal{L}}_g &= \sqrt{2} g' \sum^4_{\pi = 1} \bigg[ -{1\over 2} \overline{e}_L W^\dagger_{E_L}\widetilde{e}_L N_n(H_{n\widetilde{B}} + \cot \theta_W H_{n\widetilde{w}_{3}}) + \overline{e}^c_L W^\dagger_{E_R} \widetilde{e}_R N_n H_{n\widetilde{B}}\cr &+ {1\over 2} \cot \theta_W\overline{\nu}_L\widetilde{\nu}_L N_n H_{n\widetilde{w}_{3}}\cr &+\overline{u}_L W^\dagger_{U_L} \widetilde{u}_L N_n({1\over 6} H_{n\widetilde{B}}+{1\over 2}\cot\theta_W H_{n\widetilde{w}_3}) +\overline{d}_LW^\dagger_{D_L} \widetilde{d}_L N_n({1\over 6} H_{n\widetilde{B}}-{1\over 2}\cot\theta_WH_{n\widetilde{w}_3})\cr &- {2\over 3} \overline{u}_L^c W^\dagger_{U_R} \widetilde{u}_R N_n H_{n\widetilde{B}} + {1\over 3} \bar{d}^c_L W^\dagger_{D_R}\widetilde{d}_R N_n H_{n\widetilde{B}} + h.c.\bigg]\cr &+ g\sum^2_{c=1} [ \bar{e}_L W^\dagger_{E_L} \widetilde{\nu}_L (\chi_c K_{c\widetilde{w}}) +\bar{\nu}_L \widetilde{e}_L (\chi^\dagger_c K^*_{c\widetilde{w}})\cr &+ \bar{d}_L W^\dagger_{D_L} \widetilde{u}_{L} (\chi_c K_{c\widetilde{w}}) + \bar{u}_L W^\dagger_{U_L}\widetilde{d}_L (\chi^\dagger_c K^* _{c\widetilde{w}}) +h.c.]\cr &+\sqrt{2} g_3 [\bar{u}_L W^\dagger_{U_L}\widetilde{u}_L\widetilde{g} + \bar{d}_L W^\dagger_{D_L} \widetilde{d}_L \widetilde{g} + \bar{u}^c_L W^\dagger_{U_R} \widetilde{u}_R\widetilde{g} + \overline{d}^c_LW^\dagger_{D_R}\widetilde{d}_R\widetilde{g} + h.c.],&(3.9)\cr} $$ where\footnote{Neutrino masses are not discussed here and we choose the neutrino to be in the sneutrino mass eigenstate basis.} the neutralino and chargino mass eigenstates are related to the gauge eigenstates by e.g. $\widetilde{B} = \sum^4_{n=1} H_{n\widetilde{B}}N_n,\, \widetilde{w}_3 = \sum^4_{n=1} H_{n\widetilde{w}_3} N_n,\, \widetilde{w}^+ = \sum^2_{c=1} K_{c\widetilde{w}}\chi_c$, and $$ \eqalignno{ W_{E_L} &= U_L^\dagger V_{E_L}, \; W_{E_R} = U^\dagger_E V_{E_R}, \; W_{U_L} = U^\dagger_Q V_{U_L}, \; W_{D_L} = U^\dagger_Q V_{D_L},\cr W_{U_R} &= U^\dagger_U V_{U_R},\; W_{D_R} = U^\dagger_D V_{D_R}.\cr} $$ \noindent There are also non-diagonal chirality-changing mass terms $$ \eqalignno{ - {\cal{L}}^{n.d}_m &= \widetilde{e}_R^T W^*_{E_R}(A_E + \mu \tan\beta){\fam=9{\mathchar"7115 } }_E W^\dagger_{E_L} \widetilde{e}_L v_D + \widetilde{e}_R^T U^T_E \Delta{{\fam=9{\mathchar"7110 } }}_E U_L\widetilde{e}_L v_D\cr &+ \widetilde{d}_L^T W^*_{D_L}(A_D + \mu \tan\beta) {\fam=9{\mathchar"7115 } }_D W^\dagger_{D_R} \widetilde{d}_R v_D + \widetilde{d}^T_L U^T_Q {\Delta{\fam=9{\mathchar"7110 } }}_D U_D \widetilde{d}_R v_D\cr &+ \widetilde{u}_L^T W^*_{U_L} (A_U + \mu\cot\beta) {\fam=9{\mathchar"7115 } }_U W^\dagger_{U_R} \widetilde{u}_R v_U + \widetilde{u}_L^T U^T_Q {{\Delta\fam=9{\mathchar"7110 } }}_U U_U\widetilde{u}_Rv_U\cr &+ h.c.&(3.10)\cr} $$ The lepton flavor violating couplings are summarized in Fig.\ 2. In the rest of this section we discuss the flavor mixing matrices in the minimal supersymmetric standard model, minimal and general $SU(5)$ and $SO(10)$ models, with moderate or large $\tan\beta$. The results are summarized in Table 1. For the minimal supersymmetric standard model, the radiative corrections to the soft masses only come from the Yukawa interactions of the MSSM: $$ \eqalignno{ \Delta{\bf{m}}^2_Q &\propto {\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U +\kappa {\fam=9{\mathchar"7115 } }_D{\fam=9{\mathchar"7115 } }^\dagger_D,\cr \Delta{\bf{m}}^2_U &\propto 2{\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_U,\cr \Delta{\bf{m}}^2_D &\propto 2{\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_D,\cr \Delta{\bf{m}}^2_L &\propto {\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E,\cr \Delta{\bf{m}}^2_E &\propto 2{\fam=9{\mathchar"7115 } }_E{\fam=9{\mathchar"7115 } }^\dagger_E. &(3.11)\cr} $$ We have assumed a boundary condition on the scalar mass matrices ${\bf{m}}^2_{A} \propto I $ at $M_{PL}$, and $\kappa \neq 1$ represents the possibility that the proportionality constants are not universal. For moderate $\tan\beta, {\lambda}_t \gg {\lambda}_b$ so that the radiative corrections are dominated by $\lambda_t$. Thus one can neglect the ${\fam=9{\mathchar"7115 } }_D$ contribution and the only nontrivial mixing is $W_{D_L}$. For large $\tan\beta, {\lambda}_t$ and ${\lambda}_b$ are comparable, so ${\bf{m}}^2_Q$ will lie between ${\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U$ and ${\fam=9{\mathchar"7115 } }_D{\fam=9{\mathchar"7115 } }^\dagger_D$. Therefore both $W_{U_L}$ and $W_{D_L}$ are non-trivial. with KM matrix and one can ignore them. For the minimal $SU(5)$ model, there are only two Yukawa matrices, ${\fam=9{\mathchar"7115 } }_U = {\fam=9{\mathchar"7115 } }_{10}, {\fam=9{\mathchar"7115 } }_D = {\fam=9{\mathchar"7115 } }_E= {\fam=9{\mathchar"7115 } }_5$, and $$ \eqalignno{ \Delta{\bf{m}}^2_Q &\propto 3{\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U + 2\kappa {\fam=9{\mathchar"7115 } }_D {\fam=9{\mathchar"7115 } }^\dagger_D,\cr \Delta{\bf{m}}^2_U &\propto 3{\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_U + 2 \kappa {\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_D,\cr \Delta{\bf{m}}^2_D &\propto 4{\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_D,\cr \Delta{\bf{m}}^2_L &\propto 4{\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_D,\cr \Delta{\bf{m}}^2_E &\propto 3{\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U + 2\kappa{\fam=9{\mathchar"7115 } }_D{\fam=9{\mathchar"7115 } }^\dagger_D.&(3.12)\cr} $$ For moderate $\tan\beta, {\lambda}_t \gg {\lambda}_b$, we have non-trivial mixings for $W_{D_L}$ and $W_{E_R}$, as found in \cite{BH,BHS}. For large $\tan\beta$, ${\fam=9{\mathchar"7115 } }_D$ can not be ignored, giving non-trivial mixings for $W_{U_L}$ and $W_{U_R}$. For the minimal $SO(10)$ model considered in \cite{BHS,DH}, $$ \eqalignno{ \Delta{\bf{m}}^2_Q &\propto 5 {\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }_U^\dagger + 5 \kappa {\fam=9{\mathchar"7115 } }_D{\fam=9{\mathchar"7115 } }_D^\dagger,\cr \Delta{\bf{m}}^2_U &\propto 5 {\fam=9{\mathchar"7115 } }_U^\dagger{\fam=9{\mathchar"7115 } }_U + 5 \kappa{\fam=9{\mathchar"7115 } }_D^\dagger{\fam=9{\mathchar"7115 } }_D,\cr \Delta{\bf{m}}^2_D &\propto 5 {\fam=9{\mathchar"7115 } }_U^\dagger{\fam=9{\mathchar"7115 } }_U + 5\kappa{\fam=9{\mathchar"7115 } }_D^\dagger{\fam=9{\mathchar"7115 } }_D,\cr \Delta{\bf{m}}^2_L &\propto 5 {\fam=9{\mathchar"7115 } }_U^\dagger{\fam=9{\mathchar"7115 } }_U + 5\kappa{\fam=9{\mathchar"7115 } }_D^\dagger{\fam=9{\mathchar"7115 } }_D,\cr \Delta{\bf{m}}^2_E &\propto 5 {\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }_U^\dagger + 5\kappa{\fam=9{\mathchar"7115 } }_D{\fam=9{\mathchar"7115 } }_D^\dagger.&(3.13)\cr} $$ We have non-trivial mixings $W_{D_L}, W_{D_R}, W_{E_L}$, and $W_{E_R}$ for moderate $\tan\beta$ and non-trivial mixings for all $W$'s for large $\tan\beta$. For the general $SU(5)$ or $SO(10)$ models, defined in the last section, we get non-trivial mixings for all mixing matrices in general. However, in $SU(5)$ models with moderate $\tan\beta$, the splittings among ${\bf{m}}^2_D$ and ${\bf{m}}^2_L$ are too small (because they are generated by the small ${\fam=9{\mathchar"7115 } }_5(A)$) to give significant flavor changing effects. One might expect that the mixing in the $W_U$'s are smaller than those in the $W_D$'s because of the larger hierarchy in ${\fam=9{\mathchar"7115 } }_U$ compared with ${\fam=9{\mathchar"7115 } }_D$. However, a given $W$ is the product of a $U^\dagger$ (which diagonalizes the scalar mass matrix) and a $V$ (which diagonalizes the Yukawa matrix). Even if the mixings in $V_U$'s are smaller than those in $V_D$'s because of the larger hierarchies in ${\fam=9{\mathchar"7115 } }_U$, we do not have a general argument for the size of mixings in $U$ matrices. This is because $U$ diagonalizes (appropriate combinations of) known Yukawa matrices and unknown Yukawa matrices appearing above the GUT scale, (2.5). The mixings in $U^\dagger$ and $V$ can add up or cancel each other. Our only general expectation is that these new Yukawa matrices have similar hierarchical patterns as ${\fam=9{\mathchar"7115 } }_U$ or ${\fam=9{\mathchar"7115 } }_D$. Without a specific model, one can at most say that all non-trivial $W$'s are expected to be comparable to $V_{KM}$; the argument that the mixings in $W_U$'s should be smaller than is $W_D$'s is not valid. In the minimal models at moderate $\tan\beta$, the leading contributions to flavor changing processes, such as $\mu\to e\gamma$, involve diagrams with a virtual scalar of the third generation. Although such contributions are highly suppressed by mixing angles, they dominate because they have large violations of super-GIM\cite{GIM}: the top Yukawa coupling makes $m_{\widetilde{\tau}}$ very different from $m_{\widetilde{e}}, m_{\widetilde{\mu}}$. At large $\tan\beta$, the strange/muon Yukawa couplings get enhanced, so the splitting between $m_{\widetilde{e}}$ and $m_{\widetilde{\mu}}$ increases, leading to potentially competitive contributions to flavor changing processes which do not involve the third generation. The importance of these new diagrams can be estimated by comparing the contributions to $\Delta m^2_{21}$ (in a basis where gaugino vertices are diagonal) when the super-GIM cancellation is between scalars of the first two generations (2-1) and third generations (3-1): $$ {\Delta m^2_{21}(2\hbox{-}1)\over \Delta m^2_{21} (3\hbox{-}1)} \simeq {V_{cd}\lambda^2_2\over V_{td}V_{ts}\lambda^2_t} \simeq \left\{ \matrix{ 10^{-2}, & \hbox{for}\; \lambda_2 = \lambda_c,\cr \left({\tan\beta\over 60}\right)^2, &\hbox{for}\; \lambda_2 = \lambda_s.}\right.\eqno(3.14) $$ We can see that for large $\tan\beta$ (or any $\tan\beta$ with small $\lambda_s$ coming from the mixing of Higgs at $M_G$ i.e., $\lambda_s (M_G) = {\tan\beta\over 60} \lambda_2 (M_G))$, this could be comparable to the flavor violating effects from the large splitting of the third generation scalar masses. However, for the $\mu \to e\gamma$ in $SO(10)$ models, it does not contribute to diagrams which are proportional to $m_\tau$, (because it does not involve the third generation scalars), the dominant contributions are still those diagrams considered in \cite{BHS}. For flavor changing processes which do not need chirality flipping, such as $K-\overline{K}$ mixing, and all flavor changing processes in $SU(5)$ models, this non-degeneracy between the first two generations is important. The above discussion is summarized in Table 1. \newpage \centerline{\bf{\large Table 1}} \medskip \begin{center} \begin{tabular}{|c|c|c|c|c|c|} \hline & &\multicolumn{2}{c|}{$SU(5)$} & \multicolumn{2}{c|}{$SO(10)$}\\ \hline &MSSM&Minimal&general&minimal&general\\ \hline \hline $\delta m^2_3$&$\surd$&$\surd$&$\surd$&$\surd$&$\surd$\\ \hline $\delta m^2_2$&$\bullet$&$\bullet$&$\circ$&$\bullet$&$\circ$\\ \hline \hline $W_{U_L}$&$\bullet$&$\bullet$&$\surd$&$\bullet$&$\surd$\\ \hline $W_{D_L}$&$\surd$&$\surd$&$\surd$&$\surd$&$\surd$\\ \hline $W_{U_R}$&---&$\bullet$&$\surd$&$\bullet$&$\surd$\\ \hline $W_{D_R}$&---&---&$\surd^*$&$\surd$&$\surd$\\ \hline $W_{E_L}$&---&---&$\surd^*$&$\surd$&$\surd$\\ \hline $W_{E_R}$&---&$\surd$&$\surd$&$\surd$&$\surd$\\ \hline \end{tabular} \end{center} \medskip {\bf{Table 1:}} Summary table for the flavor mixing matrices: $ \delta m^2_3$ : important effects due to some third generation scalars not degenerate with those of first two generations. $\delta m^2_2$ : non-negligible effects due to non-degeneracy of the scalars; of the first two generations. $W_i$ : fermion $i$ and scalar $\widetilde{i}$ are rotated differently to get to mass basis. $\surd$ : present for any value of $\tan\beta$. $\bullet$ : present only for large $\tan\beta$. $\circ$ : present for large $\tan\beta$, but model dependent for moderate $\tan\beta$. --: not present. $*$ : although present, its effect for moderate $\tan\beta$ on flavor violation is small due to the small non-degeneracy among different generation scalars. \newpage \noindent{\bf{IV. Phenomenology from up-type mixing}} \medskip As discussed in the previous section, unlike the minimal models with moderate $\tan\beta$ studied in \cite{BH,BHS,DH,BHS2} in generic GUT's (for any $\tan\beta$) and even for minimal GUT's (at large $\tan\beta$), we expect mixing matrices in the up sector. Having motivated an origin for non-trivial up mixing matrices $W_{U_{L(R)}} \neq 1$, we consider some effects they produce. In the following we simply assume some $W_{u_{L(R)}}$ at the weak scale and consider their phenomenological consequences. (See however Section V and the appendix for a discussion of the scaling of mixing matrices from GUT to weak scales.) In particular we discuss $D - \bar{D}$ mixing, corrections to up-type quark masses, contributions to the neutron electric dipole moment (e.d.m.) and the possibility of different dominant proton decay modes than those expected from minimal models. \noindent{\bf IVa. $D-\bar{D}$ mixing:} To get an idea for the contribution of up-type mixing matrices to $D-\bar{D}$ mixing, we follow \cite{GM,NS} and employ the mass insertion approximation. The bounds obtained from $D-\bar{D} $ mixing on the $6\times 6$ up-squark mass matrix $m^2_U = \pmatrix{ m^2_{U_{LL}}&m^2_{U_{LR}}\cr m^2_{U_{RL}} & m^2_{U_{RR}} }$ (in the basis where gluino and Yukawa couplings are diagonal) are summarized in \cite{NS}. For average up-squark mass of $\widetilde{m} = 1 $ TeV, they are $$ \sqrt{ {m^2_{U_{LL 12}}\over \widetilde{m}^2} {m^2_{U_{RR 12}}\over \widetilde{m}^2} } \leq .04, \eqno(4.1) $$ $$ {m^2_{U_{LR 12}}\over \widetilde{m}^2} \leq .06 . \eqno(4.2) $$ Consider first (4.1). In the last section we estimated that the contribution to $m^2_{12}$ from the slight non-degeneracy between the first two generation scalars is generically at most comparabale to that from the non-degeneracy between the first two and third generation scalars. Thus, for our calculation, we only consider the contribution from the splitting between first two and third generation scalars. Then, for $A=L, R$ $$ \abs{ {m^2_{AA 12}\over \widetilde{m}^2}} = \abs{ W_{U_{A 13}} W^\dagger_{U_{A 32}}} \abs{m^2_{\widetilde{t}_A}-m^2_{\widetilde{u}_A}\over \widetilde{m}^2} \leq \abs{ W_{U_{A 13}}W^\dagger_{U_{A 32}} }.\eqno(4.3) $$ We see that for $W$'s of the same size as the corresponding KM matrix elements, the LHS of (4.1) is of order $4 \times 10^{-4}$, and the bound is easily satisfied. Turning to (4.2), note that if ${\fam=9{\mathchar"7110 } }_U = A{\fam=9{\mathchar"7115 } }_U,\; m^2_{U_{LR 12}} =0$. However, we expect ${\fam=9{\mathchar"7110 } }_U = A{\fam=9{\mathchar"7115 } }_U + \Delta{\fam=9{\mathchar"7110 } }_U$, with $\Delta{\fam=9{\mathchar"7110 } }_U$ induced in running from $M_{PL}$ to $M_G$ having primarily a third generation component in the gauge eigenstate basis. If all relevant mixing matrix elements are of order the KM matrix elements, we expect $\abs{ {m^2_{U_{LR 12}}\over \widetilde{m}^2} } = O \left(\abs{ {Am_t\over \widetilde{m}^2} V_{td}V_{ts}}\right).$ Again, we see that the bound (4.2) is generically easily satisfied, and thus we do not in general expect significant contributions to $D-\bar{D}$ mixing. \medskip \noindent{\bf IVb. Weak-scale corrections to up-type quark masses:} It is well known that there are important weak-scale radiative corrections to the down quark mass matrix proportional to $\tan\beta$ \cite{HRS,RS,BPR,COPW,H}. In general unified models, with non-zero $W_U$, there are also importrant weak scale corrections to the up quark mass matrix. {}From the diagram in Fig.\ 3, we have a contribution to up-type masses proportional to $m_t$. We find, again assuming degeneracy between the scalars of the first two generations, $$ \eqalignno{ \Delta m^{ij}_u &= {8\over 3} \left( {\alpha_s\over 4\pi}\right) m_t \left( {A +\mu \cot\beta \over M_{\widetilde{g}} }\right) W_{U_{L3i}} W_{U_{R3j}} \left[ h (x_{t_L}, x_{t_R}) - h (x_{t_L}, x_{u_R})\right.\cr & -\left. h(x_{u_L}, x_{t_R}) + h(x_{u_L}, x_{u_R})\right],&(4.4)\cr} $$ where $$ x_i \equiv {\widetilde{m}^2_i\over M^2_{\widetilde{g}}},\;\; h(x,y) = {1\over x-y} \left[ {x \log x\over 1-x} - {y\log y\over 1-y}\right].\eqno(4.5) $$ The largest fractional change in the mass occurs for the up quark. If $W_{U_{L(R)31}} $ is comparable to the corresponding KM matrix element, the contribution to ${\Delta m_u\over m_u}$ is not significant. However, if each of the $W_{U_{L(R)31}}$ are a factor 3 larger than the corresponding KM elements we can get sizable contributions. In Fig.\ 4, we plot ${\Delta m_u\over m_u}$ in ${m_{\tilde{u}}\over M_{\tilde{g}}} - {m_{\tilde{t}}\over m_{\tilde{u}}}$ space, where we have assumed $m_{\widetilde{u}_L} = m_{\widetilde{u}_R}\equiv m_{\widetilde{u}} ;\; m_{\widetilde{t}_L} = m_{\widetilde{t}_R} \equiv m_{\widetilde{t}}$, and we have put $|W_{U_{L 31}}|= |W_{U_{R 31}}| = 1/30,\; (A+ \mu \cot \beta)/m_{\widetilde{t}} =3$. Any deviations from these values can simply be multiplied in $\Delta m_u/m_u$. In some regions of the parameter space it is possible to get the entire up quark mass as a radiative effect. \newpage \noindent{\bf IVc. Neutron e.d.m.:} If we attach a photon in all possible ways to the diagram giving the contribution to $u$-quark mass, we get a contribution to the $u$-quark e.d.m., which is proportional to $m_t$ for any value of $\tan\beta$. Evaluating the diagram, we find $$ d^u = e|F|\sin\phi_u\eqno(4.6) $$ where $$ \eqalignno{ F= &{8\over 3}\left({\alpha_s \over 4{\pi} }\right) m_t {A + \mu \cot \beta\over M^3_{\widetilde{g}} } W_{U_{L31}} W_{U_{L33}}^*W_{U_{R31}} W^*_{U_{R33}}\cr &\times \bigg[ \widetilde{G}_2 (x_{t_{L}}, x_{t_{R}}) - \widetilde{G}_2(x_{t_{L}}, x_{u_{R}}) -\widetilde{G}_2 (x_{t_{R}}, x_{u_{L}}) + \widetilde{G}(x_{u_{L}}, x_{u_{R}} )\bigg],&(4.7) \cr} $$ where $$\widetilde{G}_2 (x, y) = {g(x) - g(y)\over x-y},\; \; g(x) = {1\over 2(x-1)^3} [x^2-1-2x \log x]\eqno(4.8) $$ and $$ Im \left[ m_t W_{U_{L31}} W_{U_{L33}}^* W_{U_{R31}} W^*_{U_{R33}} \right] \equiv | m_t W_{U_{L31}}W^*_{U_{R33}}W_{U_{R31}} W^*_{U_{R33}}|\sin \phi_u.\eqno(4.9) $$ In general we expect a large non-zero sin $\phi_u$. If the combination of $W$'s appearing in the above is comparable to the combination giving a down quark e.d.m., the $u$-quark contribution will dominate over the $d$-quark contribution to the neutron e.d.m.\ considered in \cite{DH} by a factor ${m_t\over 4m_b\tan\beta}$, (the factor 4 comes from the quark model result $d_n=4/3 d_d - 1/3 d_u$). Hence, the neutron e.d.m.\ may be competitive with $\mu \to e\gamma$ and $d_e$ as the most promising flavor changing signal for supersymmetric unification. \noindent{\bf IVd. Proton decay:} Finally we turn briefly to the relevance of up-type mixing matrices for proton decay; in particular to the important question of the charge of the lepton in the final state. We know that upon integrating out the superheavy Higgs triplets we can generate the baryon number violating operators ${1\over 2M_H} (QQ)(QL)$ and ${1\over M_H} (EU)(DU)$ in the superpotential. These operators must subsequently be dressed at the weak scale in order to obtain four-fermion operators leading to proton decay. The dressing may be done with neutralinos, charginos or gluinos where possible. Since the dressed operator grows with gauge couplings and vanishes for vanishing neutralino/chargino/gluino mass, one might naively expect gluino dressing to be most important. However, if the up-type mixing matrices are trivial, gluino dressed operators can only lead to proton decay with a neutrino in the final state. To see this, we examine each operator separately: $(eu_a)(d_bu_c)\epsilon^{abc}$ (where $a,b,c$ are color indices) must involve $u$'s from two different generations because of the $\epsilon^{abc}$. One of them has to be a $u$, so the other is a $c$ or a $t$. If there is no up mixing, the up flavor does not change in the dressing process, so the final state would have to contain a $c$ or a $t$. Since $m_t, m_c > m_p$, this can not happen. Next, consider $(QQ)(QL) = u^a_Ld^b_L(u^c_Le_L-d^c_L v_L)\epsilon_{abc}$. By exactly the same argument as the above, the $u_L^a d^b_L u^c_Le_L\epsilon_{abc}$ operator can not contribute to proton decay. Thus, we see that in the absence of mixing in the up sector, gluino dressing can only give neutrinos in the final state. However, the above arguments break down if up-mixing matrices are non-trivial, since gluino dressed diagrams give a significant contribution to the branching ratio for charged lepton modes in proton decay. A detailed study of flavor mixing in the up sector \cite{A} concludes that, whether the wino or gluino dressings are dominant, the muon final state in proton decay is of greatly enhanced importance. Without the mixings, one expects ${\Gamma (p\to K^o \mu^+) \over \Gamma (p\to K^+\bar{\nu})} \approx 10^{-3}$. The up mixing in general models increases this by $O(100)$ making the mode $p\to K^o\mu^+$ a favorable one for discovery of proton decay. \newpage \noindent{\bf Section V. Large $\tan\beta$: Analytic Treatment} \medskip The large $\tan\beta$ scenario is interesting for a number of reasons. For moderate $\tan\beta$, the only way to understand $m_t \gg m_b, m_\tau$ is to have $\lambda_t \gg \lambda_b, \lambda_\tau$ at the weak scale. This gives us little hope of attributing a common origin for third generation Yukawa couplings at a higher scale. However, for large $\tan\beta \sim {{O}}\left({m_t\over m_b}\right)$, the weak scale $\lambda_t, \lambda_b, \lambda_\tau$ are comparable and the above hope is restored. (In fact it is realized in $SO(10)$ models like the ADHRS example outlined in section VII). For us, this is sufficient motivation to study the large $\tan\beta$ case in more detail. Also, this case was not studied in \cite{BHS}. We shall see that unexpected new features arise in the large $\tan\beta$ limit. The largest contribution to the $\mu \to e\gamma$ amplitude comes from the diagram with $L-R$ scalar mass insertion (Fig.\ 5). In the $L-R$ insertion approximation, the amplitude for $\mu_{L(R)}$ decay is $$ \eqalignno{ F_{L(R)} &= {\alpha\over 4\pi\cos^2\theta_W} m_\tau W_{E_{L(R)_{32}}} W_{E_{R(L)_{31}}} W^*_{E_{L(R)_{33}}} W^*_{E_{R(L)_{33}}} (A_E + \mu \tan\beta)\cr &\times \left[ G_2(m^2_{\widetilde{\tau}_L}, m^2_{\widetilde{\tau}_R}) - G_2(m^2_{\widetilde{e}_L}, m^2_{\widetilde{\tau}_R}) -G_2 (m^2_{\widetilde{\tau}_L}, m^2_{\widetilde{e}_R})+ G_2 (m^2_{\widetilde{e}_L }, m^2_{\widetilde{e}_R})\right],\cr} $$ where $$ \eqalignno{ G_2(m^2_1, m^2_2) &= {G_2(m^2_1) - G_2(m^2_2)\over m^2_1 - m^2_2},\cr G_2 (m^2) &= \sum^4_{n =1} (H_{n\widetilde{B}} + \cot \theta_W H_{n\widetilde{w}_3}) g_2\left( {m^2\over M^2_n}\right).&(5.1)\cr} $$ Note, however, that for large $\tan\beta$ the $L-R$ insertion approximation may be a bad one, since the chirality changing mass for the third generation becomes comparable to the chirality conserving masses. A correct treatment will be used for the numerical analysis in the next section. We still expect, however, that the amplitude to be proportional to $W_{E_{32}}W_{E_{31}}$ because of the unitarity of the mixing matrices: the sum of contributions from the first two generations is proportional to $W_{1i}W^*_{1j} + W_{2i} W^*_{2j} =- W_{3i}W^*_{3j}$ for $i\neq j$, and the contribution from the third generation is itself proportional to $W_{3i}W^*_{3j}$. Two simplifications in the dependence of the $\mu\to e\gamma$ rate on parameter space occur for large $\tan\beta$. First, since the dominant diagram involves the $L-R$ insertion $(A + \mu\tan\beta)m_t$, and since $\tan\beta$ is large, the amplitude does not depend on the weak scale parameter $A$. Second, in the large tan$\beta$ limit, the chargino mass matrix is $$ M_\chi = \pmatrix{ M_2 & \sqrt{2} M_W \sin \beta\cr \sqrt{2} M_W \cos\beta &-\mu} \longrightarrow \pmatrix{ M_2 & \sqrt{2}M_W\cr 0& -\mu},\eqno(5.2) $$ and the parameters $M_2, \mu$ have a direct interpretation as the chargino masses. (Note that this assures us that $\mu\tan\beta$ will likely always be much bigger than $A$; for a $\tan\beta$ of 50, the LEP lower bound on chargino mass of 45 GeV tells us that $\mu \tan\beta > 2 $ TeV, so for $A$ to be comparable to $\mu\tan\beta$ we must have $A > 2$ TeV.) In considering $\mu \to e\gamma$ for large $\tan\beta$, two factors come immediately to mind which tend to (perhaps dangerously) enhance the rate over the case with moderate $\tan\beta$. (i) As we have already mentioned, the dominant contribution to $\mu \to e\gamma$ grows with $\tan\beta$; the diagram in Fig.\ 5 is proportional to $\tan\beta$, a factor of 900 in the rate for $\tan\beta = 60$ compared to $\tan\beta =2$. (ii) For large $\tan\beta,\; \lambda_\tau$ can be $O(1)$ and we can not neglect its contribution to the running of the slepton mass matrix from $M_G$ to $M_S$ (soft SUSY breaking scale). This scaling generally splits the third generation slepton mass even further from the first two generations, meaning a less effective super-GIM mechanism and a larger amplitude for $\mu \to e\gamma$. While both of the above effects certainly exist, there are also two sources of {\it{suppression}} of the amplitude for large $\tan\beta$, which can together largely compensate for the above factors: (i)$'$ Large $\tan\beta$ allows $\lambda_t$ to be smaller than for moderate tan$\beta$. There are two reasons for this. First, large $\tan\beta$ allows $v_U$ to be larger and so $\lambda_t$ can be smaller to reproduce the top mass. Secondly, $b-\tau$ unification \cite{CEG} is achieved with a smaller $\lambda_\tau$ in the large $\tan\beta$ regime \cite{HRS,RS}. Since $\lambda_t$ is smaller, a smaller non-degeneracy between the third and first two generations is induced in running from $M_{PL} $ to $M_G$, suppressing the amplitude compared to the moderate $\tan\beta$ case. (ii)$'$ In comparing large and moderate tan$\beta$, we must know how the mixing matrices $W_{L,R_{3i}}$ (appearing at the vertices of the diagrams responsible for $\mu\to e\gamma$) compare in these two cases. In the moderate tan$\beta$ minimal models discussed in \cite{BHS}, $W_{L,R_{3i}}$ were equal to the corresponding KM matrix elements $V_{KM3i}$ at $M_G$, and this equality was approximately maintained in running form $M_G$ to $M_S$. As discussed in the previous sections, for more general models one expects that the $W_{L(R)_{3i}}$ at $M_G$ are equal to $V_{KM3i}$ at $M_G$ up to some combination of Clebsches. One might then expect (as in the minimal models) that this relationship continues to approximately hold at lower scales. In fact for large $\tan\beta$ this expectation is false. We find that often, the $W_{L(R)3i}$ {\it{decrease}} from $M_G$ to $M_S$, overcompensating for the increased non-degeneracy between the third and first two generation slepton masses induced by large $\lambda_\tau$ (point (ii) above). In the following, we examine the scaling of these mixing matrices in detail. Consider first the lepton sector. The renormalization group equation (RGE) for ${\fam=9{\mathchar"7115 } }_E$ (in the following $t ={log \mu\over 16 \pi^2}$) is $$ - {d{\fam=9{\mathchar"7115 } }_E\over dt} = {\fam=9{\mathchar"7115 } }_E [3{\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E + Tr (3{\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_D + {\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E) - 3 g^2_2 - {9\over 5} \; g^2_1]\eqno(5.3) $$ giving $$ - {d\over dt}{\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E = 6 {\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E + 2 {\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E Tr (3{\fam=9{\mathchar"7115 } }^\dagger_D {\fam=9{\mathchar"7115 } }_D + {\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E) - (6g^2_2+{18\over5} \; g^2_1) {\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E \eqno(5.4) $$ $$ - {d\over dt} {\fam=9{\mathchar"7115 } }_E {\fam=9{\mathchar"7115 } }_E^\dagger = 6 {\fam=9{\mathchar"7115 } }_E{\fam=9{\mathchar"7115 } }_E^\dagger + 2 {\fam=9{\mathchar"7115 } }_E{\fam=9{\mathchar"7115 } }^\dagger_E Tr (3{\fam=9{\mathchar"7115 } }^\dagger_D {\fam=9{\mathchar"7115 } }_D + {\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E) - (6g^2_2 + {18\over 5}\; g^2_1) {\fam=9{\mathchar"7115 } }_E {\fam=9{\mathchar"7115 } }_E^\dagger . \eqno(5.5) $$ These in turn imply that the basis in which ${\fam=9{\mathchar"7115 } }^\dagger_{{E}} {\fam=9{\mathchar"7115 } }_{{E}}$ is diagonal, and the (in general different) basis where ${\fam=9{\mathchar"7115 } }_{{E}}{\fam=9{\mathchar"7115 } }_{{E}}^\dagger$ is diagonal, do not change with scale. Consider now the evolution of the left handed slepton mass matrix ${\bf{m}}^2_L$. The RGE for ${\bf{m}}^2_L$ is $$ -{d\over dt} {\bf{m}}^2_L = ({\bf{m}}^2_L + 2m^2_{H_d}){\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E + 2 {\fam=9{\mathchar"7115 } }^\dagger_E {\bf{m}}^2_E {\fam=9{\mathchar"7115 } }_E + {\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E {\bf{m}}^2_L + 2 {\fam=9{\mathchar"7110 } }_E^\dagger {\fam=9{\mathchar"7110 } }_E + \; \hbox{Gaugino terms} .\eqno(5.6) $$ In the basis where ${\fam=9{\mathchar"7115 } }^\dagger_{{E}} {\fam=9{\mathchar"7115 } }_{{E}}$ is diagonal, keeping only the $\lambda_\tau$ contribution, the $3i$ entry $(i\neq 3)$ becomes: $$ - {d\over dt} m^2_{L3i} = \lambda^2_{\tau} m^2_{L3i} + 2 ({\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E)_{3i} .\eqno(5.7) $$ In this basis, we have ${\bf{m}}^2_L = W^\dagger_L\overline{\bf{m}}_L^2W_L$. (Here and in the remainder of this section, we abbreviate $W_{E_{L(R)}}\to W_{L(R)}$). Assuming degeneracy between scalars of the first two generations, $ m^2_{L_{3i}} = W_{L 3i}W_{L33}^\dagger (m^2_{\tau_{L}}- m^2_{e_{L}}) \equiv W_{L3i}W^\dagger_{L 33} \Delta m^2_L. $ Then (5.7) becomes $$ - {d\over dt} (W_{L3i} W_{L33}^\dagger\Delta m^2_L) = \lambda^2_{\tau} (W_{L3i}W^\dagger_{L 33} \Delta m_L^2) + 2({\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E)_{3i} .\eqno(5.8) $$ For now, we ignore the $({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i}$ term in (5.8), yielding the solution: $$ (W_{L3i}W_{L33}^\dagger \Delta m^2_L)(M_S) = e^{-I_\tau} (W_{L3i} W_{L33}^\dagger \Delta m^2) (M_G), \eqno(5.9) $$ where $$ I_i \equiv \int_0^{\log {M_G\over M_S}} {dt\over 16\pi^2} \lambda^2_i (t). \eqno(5.10) $$ Thus, $$ W_{L 3i} W^\dagger_{L33}(M_S) = e^{-I_\tau} {\Delta m^2_R(M_G)\over \Delta m^2_R (M_S)} W^\dagger_{L3i}W_{L_{33}} (M_G).\eqno(5.11) $$ Similarly, we find $$ W_{R 3i} W^\dagger_{R33} (M_S) = e^{-2I_\tau} {\Delta m^2_R (M_G)\over \Delta m^2_R (M_S)} W_{R 3i}W^\dagger_{R_{33}} (M_G).\eqno(5.12) $$ Note that, generically the quantities ${\Delta m^2_{L(R)}(M_G)\over \Delta m^2_{L(R)} (M_S)}$ are smaller than one, since the third generation mass gets split even further from the first two generations in running from $M_G$ to $M_S$. Thus, we find that the $W_{L(R)_{3i}}$ get smaller in magnitude as we scale from $M_G$ to $M_S$, in contrast with the KM matrix elements $V_{KM3i}$, which scale as $$ V_{KM3i} (M_S) = e^{(I_t+I_b)} V_{KM 3i} (M_G).\eqno(5.13) $$ Suppose that at $M_G$ the $W_{L(R)}$ are related to $V_{KM}$ though some combination of Clebsches determined by the physics above the GUT scale. $$ W^\dagger_{L(R)33}W_{L(R)3i} (M_G) = z_{i{L(R)}} V_{KM 3i} (M_G) .\eqno(5.14) $$ This relationship is not maintained at lower scales; instead we have: $$ \hspace{-.2in} W^\dagger_{L33}W_{L 3i} (M_S) = {\Delta m^2_L(M_G)\over \Delta m^2_L (M_S)} e^{-(I_\tau+I_t+I_b)} z_{i_{L}}V_{KM 3i}(M_S), \eqno(5.15) $$ $$ W^\dagger_{R33}W_{R 3i} (M_S) = {\Delta m^2_R(M_G)\over \Delta m^2_R (M_S)} e^{-(2I_\tau + I_t+I_b)} z_{i_{R}}V_{KM 3i}(M_S).\eqno(5.16) $$ The dominant contribution to the $\mu \to e\gamma$ rate is proportional to \newline $| W^\dagger_{L33} W_{L{32}}W^\dagger_{R33} W_{R{31}} (M_S)|^2 + | W^\dagger_{L33} W_{L{31}}W^\dagger_{R33} W_{R{32}}(M_S)|^2$, giving $$ \eqalignno{ Br( \mu\to e\gamma) &= \left[ {\Delta m^2_L (M_G)\over \Delta m^2_L (M_S)} {\Delta m^2_R (M_G)\over \Delta m^2_R (M_S)} \right]^2 e^{-(6I_\tau + 4I_t + 4I_b)} \times (|z_{2_L}z_{1_R}|^2 + |z_{1_L}z_{2_R}|^2) \cr &\times Br(\mu\to e\gamma, W_{L(R)3i} W^\dagger_{L(R)33}(M_S)\to V_{KM3i}(M_S))\cr &\hspace{-.7in} \equiv \epsilon (|z_{2_L} z_{1_R}|^2 + | z_{1_L} z_{2_R}|^2) \times Br(\mu\to e\gamma, W^\dagger_{L(R)33} W_{L(R)33}(M_S)\to V_{KM3i}(M_S)).\cr &&(5.17) \cr} $$ This $\epsilon$ represents a possibly significant suppression of the rate for large $\tan\beta$. At this point, the reader may object: it is true that the $W_{L(R)3i}$ decrease from $M_G$ to $M_S$, but as already mentioned, the non-degeneracy between the third and first two generating is increasing. Which effect wins? We argue that in general there is a net suppression. This is easiest to see if in computing the $\mu \to e\gamma$ amplitude, we use the mass insertion approximation rather than mixing matrices at the the vertices (Fig.\ 6). Although this may be a poor approximation, it serves to illustrate our point. (Of course no such approximation is made in our numerical work.) {}From the diagram it is clear that the amplitude is proportional to $m^2_{L32} m^2_{R31} (M_S)$. {}From (5.7), we see that the rate scales as $$ \left( m^2_{L32}m^2_{R31}\right)^2 (M_S) = e^{-(6I_\tau +4I_t+4I_b)} \left(m^2_{L_{32}}m^2_{R_{31}}\right) (M_G),\eqno(5.18) $$ a net suppression. In the mass insertion approximation, then, the terms ${\Delta m^2(M_G)\over \Delta m^2(M_S)}$ in (5.17) serve to exactly compensate for the increased non-degeneracy between $m^2_{\widetilde{e}_L}$ and $m^2_{\widetilde{\tau}_L}$; what remains is still a suppression. This, together with (i)$'$ invalidates the naive expectation that the theory is ruled out in most regions of parameter space due to the enhancing factors (i) and (ii), (although there are still stringent constraints on the parameter space). The above analysis suggests that individual lepton number conservation is an infrared fixed point of the MSSM (whereas individual quark number conservation is an ultraviolet fixed point). A more complete analysis of scaling for the lepton sector and a discussion of scaling in the quark sector is presented in the appendix. \newpage \noindent{\bf Section VI. Large $\tan\beta$: Numerical Results} \medskip The amplitude for $\mu \to e\gamma$ depends on the $6\times 6$ slepton mass matrix ${\bf{M}}^2$. In the basis where ${\bf{m}}^2_L,{\bf{ m}}^2_E$ are diagonal, we have $$ {\bf{M}}^2 = \pmatrix{ {\bf{\overline{m}}}^2_{{E}_{L}} + {\bf{D}}_L & {\bf{k}}\cr {\bf{k}}^\dagger &{\bf{\overline{m}}}^2_{{E}_{R}} + {\bf{D}}_R}\eqno(6.1) $$ where in the large $\tan\beta$ limit, $D_i = - (T_{3i} - Q_i \sin^2\theta_W)M^2_Z$ is the $D$-term contribution, and $k_{ij} = \mu m_\tau \tan\beta W_{L 3i}W_{R 3j}$. The amplitude from Fig.\ 1 for $\mu_L$ decay is $$ F_L = {\alpha\over 4\pi\cos^2\theta_W} W^\dagger_{L i2} G_2({\bf{M}}^2)_{LR ij} W^\dagger_{Rj1} $$ where $$ G_2 ({\bf{M}}^2) \equiv \pmatrix{G_2(M^2)_{LL} & G_2(M^2)_{LR}\cr G_2(M^2)_{RL} & G_2(M^2)_{RR} }, \eqno(6.3) $$ In \cite{BHS}, ${\bf{M}}^2$ was approximately diagonalized by the $\mu m_\tau \tan\beta$ insertion approximation, and $G_2 ({\bf{M}}^2)$ was calculated using this approximate diagonalization. Since here $\tan\beta$ is large, we wish to avoid making such an approximation, and numerically diagonalize the full 6$\times 6\; {\bf{M}}^2$. Faced with a rather large parameter space, we must decide which parameters to use in our numerical work. We have firstly decided to do our analysis only for large $\tan\beta$, since the moderate $\tan\beta$ scenario has been covered in \cite{BHS}. Secondly, we choose to present our results in a different way than in \cite{BHS}, where the rates for $\mu \to e\gamma$ were plotted against a combination of Planck scale and weak scale parameters. In our work, we compute $\mu \to e\gamma$ entirely in terms of weak scale parameters. In particular, we assume that the necessary condition for a significant $\mu\to e\gamma$ rate exist at the weak scale, namely non-trivial mixing matrix $W_{L,R_{3i}}$ and non-degeneracy between third and first two generation slepton masses. In the previous sections, we have shown a possible way in which these ingredients may be produced. Our plots for $\mu \to e \gamma$ rates are made against low energy parameters, and we separately plot the regions in low energy parameter space predicted by our particular scenario for generating $\mu\to e \gamma$. This way, our plots are in terms of experimentally accessible quantities and can be thought of as constraining the parameter space of the effective 3-2-1 softly broken supersymmetric theory resulting from the spontaneous breakdown of a GUT. (We use the GUT to relate weak scale gaugino masses.) Our low energy plots have no dependence on the physics above the GUT scale, all the model dependence comes into the predictions for low energy parameters the GUT makes. If the predicted region of low energy parameters corresponds to a $\mu \to e\gamma$ rate exceeding experimental bounds, the theory is ruled out. There is a more practical reason for working directly with low-energy parameters specific to large $\tan\beta$: the well known difficulty in achieving electroweak symmetry breaking in this regime. Working with high energy parameters, and imposing universal scalar masses necessitates a fine-tune to achieve $SU(2) \times U(1)$ breaking. However, we have nowhere in our analysis made the assumption of universal scalar masses, hence the Higgs masses and squark/slepton masses are independent in our analysis, and therefore the $\mu$ parameter is not tightly constrained by squark/slepton masses. Working with weak scale parameters allows us to assume that the desired breaking has occurred without having to know the details of the breaking. With the aforementioned assumption about the existence of a GUT, and assuming degeneracy between the first two generations, the rate for $\mu \to e\gamma$ depends on the weak scale parameters $\mu, \tan\beta, M_2, m^2_{\widetilde{e}_L}, m^2_{\widetilde{\tau}_L}, m^2_{\widetilde{e}_R}, m^2_{\widetilde{\tau}_R}, $ $W_{L 3i}, W_{R 3i}$. We know that the amplitude depends on $W_{L(R)3i} $ simply through the product $W_{L 3i} W_{R 3j}$, so for normalization in our plots we put $W_{L(R)3i} = V_{KM 3i}$. Any deviation from this can be simply multiplied into the rate. We also fix $\tan\beta = 60$, and put $m_{\widetilde{\tau}_{L(R)}} = m_{\widetilde{e}_{L(R)}} - \Delta_{L(R)}$. Next, we use some high energy bias to relate $m_{\widetilde{e}_L}$ and $m_{\widetilde{e}_R}$: we assume that their difference is proportional to $M_2$ ( as would be the case if they started out degenerate and were split only through different gauge interactions), so we put $m_{\widetilde{e}_L}= m_{\widetilde{e}_R} - rM_2$. In all specific models we have looked at, $r$ is small (less than about .2). We find that, as long as $r$ is small, the rate has little dependence on its exact value, so we put $r=0, m_{\widetilde{e}_L} = m_{\widetilde{e}_R} \equiv \overline{m}_{\widetilde{e}}$. We also found that as long as ${\Delta_L\over \Delta_R}$ is close to 1, there is little dependence on its actual value either, so we put $\Delta_L=\Delta_R\equiv \Delta$. Now, the $\mu \to e\gamma$ rate depends only on $\mu, M_2, \overline{m}_{\widetilde{e}}$ and $\Delta$, and we have the large $\tan\beta$ interpretation of $\mu$ and $M_2$ as chargino masses. Fixing $\overline{m}_{\widetilde{e}}=300$ GeV, we make contour plots of $Br(\mu \to e\gamma)$. The rate scales roughly as ${\overline{m}_{\widetilde{e}}}^{-4}$ and $\mu^2$ for scalar masses heavy compared with gaugino masses. In Fig.\ 7, we fix $\mu$ and plot in $M_2-\Delta$ space. In Fig.\ 8, we fix $\Delta$ and plot in $\mu - M_2$ space. In Fig.\ 9, we plot the values of $\Delta$ predicted by the GUT against $M_2$, for various values of $\lambda_t (M_G)$ and $A_e(M_S)$ and for two values of $b_5$, the gauge beta function coefficient above the GUT scale. In Fig.\ 10, we plot the suppression factor $\epsilon$ for the same parameter set as in Fig.\ 9. We see that, over a significant region in parameter space, $\epsilon$ is small, between 0.2 and 0.01. It is clear from Fig.\ 7 that, with no suppression, a typical value for $\Delta$ of 0.3 ($\times$300GeV) would give rise to rates above the current bound of $Br(\mu \to e\gamma) < 4.9 \times 10^{-11}$\cite{B}. However, from Fig.\ 10, the suppression from $\epsilon$ is seen to be typically 20, allowing $ \Delta$'s of up to 0.45 ($\times$300GeV). We see that $\epsilon $ is crucial in giving the GUT more breathing room, as $\Delta$'s of less than 0.45 are more common. {}From Fig.\ 8 it is also clear that regions of small $\mu$ and $M_2$ (that is, light chargino masses) are preferred. Smaller $\mu$ is preferred because it decreases the L-R mass $\mu m_\tau \tan\beta$, small $M_2$ is preferred because in the limit that the neutralino mass tends to zero, the diagrams Fig.\ 5 vanish. We also note that smaller $\mu, M_2$ are preferred for electroweak symmetry breaking\cite{HRS,RS}. If $\mu$ and $M_2$ are both small, the lightest supersymmetric particle (LSP) can be quite light, (but where it has significant higgsino component, it must be heavier than 45 GeV in order to be consistent with the precise measurement of the $Z$ width), and it annihilates (primarily through its higgsino components) through a $Z$ into fermion antifermion pairs much like a heavy neutrino. The contribution of the LSP to energy density of the universe $\Omega h^2$ then just depends on its mass, and the size of its higgsino components, both of which only depend on $\mu$ and $M_2$ in the large $\tan\beta$ limit. In Fig.\ 11, we make a plot of $\Omega h^2$ in $\mu-M_2$ space. We see that it is possible to get $\Omega \sim O(1)$ in some region of the parameter space. \newpage \noindent{\bf{VII. The Example of ADHRS Models}} \medskip In this section, we study the ADHRS models \cite{ADHRS} which are known to give realistic fermion masses and mixing patterns. These models are specific enough for us to do calculations and make some real predictions. Although not necessarily correct, they are good representatives of general GUT models. We believe that by studying them, one can see in detail the general features of generic realistic GUT models and the differences between them and the minimal $SU(5)$ or $SO(10)$ models. As mentioned in Sec.\ II, in ADHRS models, the three families of quarks and leptons lie in three 16 dimensional representations of $SO(10)$, and the two low energy Higgs doublets lie in a single 10 dimensional representation. Only the third generation Yukawa couplings come from a renormalizable interaction $$ \lambda_{33} 16_3 16_3 10 .\eqno(7.1) $$ All other small Yukawa couplings come from nonrenormalizable interactions after integrating out the heavy fields. These interactions can be written in general as $$ 16_i \lambda_{ij} (A_a) 16_j 10.\eqno(7.2) $$ The $A_{a}$'s are fields in the adjoint representation of $SO(10)$ and their vevs break $SO(10)$ down to the Standard Model gauge group. Therefore, these Yukawa couplings can take different values for fermions of the same generation with different quantum numbers under $SU(3)\times SU(2) \times U(1)$ and a realistic fermion mass pattern and nontrivial KM matrix can be generated. In ADHRS models, the minimal number (four) of operators is assumed to generate the up, down-type quark and charged lepton Yukawa coupling matrices ${\fam=9{\mathchar"7115 } }_U, {\fam=9{\mathchar"7115 } }_D$ and ${\fam=9{\mathchar"7115 } }_E$, and they take the form at $M_G$ $$ {\fam=9{\mathchar"7115 } }_U = \pmatrix{ 0 &z_uC &0\cr z'_uC&y_u E & x_uB\cr 0 & x'_uB& A}, {\fam=9{\mathchar"7115 } }_D = \pmatrix{ 0 &z_dC &0\cr z'_dC & y_dE & x_dB\cr 0 & x'_dB & A}, {\fam=9{\mathchar"7115 } }_E=\pmatrix{ 0 & z_eC & 0\cr z'_eC& y_eE & x_eB\cr 0 & x'_eB & A}, \eqno(7.3) $$ where the $x, y, z$'s are Clebsch factors arising from the VEV's of the adjoint Higgs fields $A_a$. This form is known to give the successful relations $V_{ub}/V_{cb} = \sqrt{m_u/m_c}$ and $V_{td}/V_{ts} = \sqrt{ m_d/m_s}$ \cite{HR} so it is well motivated. Strictly speaking, the interaction (7.2) become the usual Yukawa form only after the adjoints $A_a$ take their VEV's at the GUT scale. However, as we explained in Sec.\ II, they can be treated as the usual Yukawa interactions up to the ultraheavy scale (which we will assume to be $M_{PL}$) where the ultraheavy fields are integrated out if the wavefunction renormalizations of $A_a$'s are ignored. In the one-loop approximation which we use later in calculating radiative corrections from $M_{PL}$ to $M_G$, they give the same results, because the wavefunction renormalizations of the adjoints $A_a$ only contribute at the two-loop order. This makes our analysis much easier. Above the GUT scale, in addition to the Yukawa interactions (2.4) which give the fermion masses, we have the interactions (2.5) as well. Each Yukawa matrix has different Clebsch factors $x, y, z$ associated with its elements. All the Yukawa matrices have the ADHRS form $$ \lambda_I = \pmatrix{ 0& z_I C & 0\cr z'_I C & y_IE & x_IB\cr 0 & x'_I B & A },\; I = qq, eu, ud, q\ell, nd, n\ell .\eqno(7.4) $$ If each entry of the Yukawa matrices is generated dominantly by a single operator, like in the ADHRS models, then the phases of the same entries of all Yukawa matrices are identical. One can remove all but the $\lambda_{22}$ phases by rephasing the operators. After phase redefinition only $E$ is complex and is responsible for CP violation. In order to generate the realistic fermion mass and mixing pattern, one expects the following hierarchies, $$ \eqalignno{ {B\over A} &\sim V_{cb} \hskip .14in\sim \epsilon^2, \cr {E\over A} &\sim {m_s\over m_b} \hskip 8pt\sim \epsilon^2, \cr {C\over E} &\sim \sin \theta_c \sim \epsilon, \;\;\; \hbox{where } \; \epsilon \sim 0.2.&(7.5)\cr} $$ The hierarchical Yukawa matrices can be diagonalized approximately \cite{HR}, the unitary rotation matrices which diagonalize them at the GUT scale can be approximately written as $$ {\fam=9{\mathchar"7115 } } =\pmatrix{ 0& zC & 0\cr z'C & y\abs{E}e^{\widetilde{\phi}}&xB\cr 0 & x'B&A } = V^*_F \bar{\fam=9{\mathchar"7115 } } V^\dagger_B,\eqno(7.6) $$ $$ V_F \simeq \pmatrix{ e^{i\phi} & S_{F_1} e^{i\phi} & 0\cr -S_{F_1} & 1 & S_{F_2}\cr S_{F_1}S_{F_2} & -S_{F_2} & 1\cr},\eqno(7.7) $$ $$ V_B \simeq \pmatrix{ 1 & S_{B_1} & 0\cr -S_{B_1} e^{-i\phi} & e^{-i\phi}& S_{B_2} \cr S_{B_1} S_{B_2} e^{-i\phi} & -S_{B_2}e^{-i\phi} & 1},\eqno(7.8) $$ where $$ \eqalignno{ S_{F_2} &= {xB \over A},\; S_{B_2} = {x'B \over A},\cr S_{F_1} &= {zC \over E'},\; S_{B_1} = { z'C \over E'},\cr E' &= \abs{ y E - S_{F2} S_{B2} A } = \abs{ y |E| e^{i\widetilde{\phi} } - {x'x B^2\over A} },\cr \widetilde{\phi} &= arg (E),\; \phi = arg (y E - {x'x B^2\over A} ). } $$ The soft SUSY-breaking scalar masses for the three low energy generations and trilinear $A$ terms are assumed to be universal\footnote{If the nonrenormalizable operators already appear in the superpotential of the underlying supergravity theory, the $A$ terms will be different for different dimensional operators, and will induce unacceptably large $\mu \to e\gamma$ rate because the triscalar interactions and the Yukawa interactions can not be diagonalized in the same basis for the first two generations. In theories where the nonremormalizable operators come from integrating out heavy fields at $M_{PL}$ and all the relevant interactions have the same $A$ term, the resulting nonrenormalizable operators will also have the same $A$ term.} at Planck scale $M_{PL}$ as in \cite{BHS}. Beneath $M_{PL}$, the radiative corrections from the Yukawa couplings destroy the universalities and render the mixing matrices non-trivial. In the one-loop approximation, the radiative corrections to the soft SUSY-breaking parameters at $M_{G}$ are simply related to the Yukawa coupling matrices and therefore the relations between general mixing matrix elements and KM matrix elements are also simple. This allows us to see the similar hierarchies in the general mixing matrices and the KM matrix very clearly. Although the one-loop approximation may not be a good approximation for quantities involving third generation Yukawa couplings, we will be satisfied with it since it simplifies things a lot and the uncertainties in other quantities such as Clebsch factors are probably much bigger than the errors made in the one-loop approximation. The RG equations, for ${\bf{m}}^2_E$ as an example, from $M_{PL}$ to $M_G$ are $$ \eqalignno{ {d\over dt} {\bf{m}}^2_E &= {1\over 16\pi^2} \bigg[2(2{\fam=9{\mathchar"7115 } }_E{\bf{m}}^2_L {\fam=9{\mathchar"7115 } }^\dagger_E + 2 {\fam=9{\mathchar"7115 } }_E m^2_{H_D} {\fam=9{\mathchar"7115 } }^\dagger_E + {\bf{m}}^2_E {\fam=9{\mathchar"7115 } }_E {\fam=9{\mathchar"7115 } }^\dagger_E + {\fam=9{\mathchar"7115 } }_E{\fam=9{\mathchar"7115 } }_E^\dagger {\bf{m}}^2_E + 2 {\fam=9{\mathchar"7110 } }_E{\fam=9{\mathchar"7110 } }^\dagger_E)\cr &+ 3(2 {\fam=9{\mathchar"7115 } }_{eu} {\bf{m}}^2_U {\fam=9{\mathchar"7115 } }^\dagger_{eu} + 2 {\fam=9{\mathchar"7115 } }_{eu} m^2_{H_{U_3}} {\fam=9{\mathchar"7115 } }^\dagger_{eu} + {\bf{m}}^2_E {\fam=9{\mathchar"7115 } }_{eu} {\fam=9{\mathchar"7115 } }^\dagger_{eu} + {\fam=9{\mathchar"7115 } }_{eu} {\fam=9{\mathchar"7115 } }^\dagger_{eu} {\bf{m}}^2_E + 2{\fam=9{\mathchar"7110 } }_{eu} {\fam=9{\mathchar"7110 } }^\dagger_{eu})\cr &- \hbox{gaugino mass contribution}\bigg].&(7.9)\cr} $$ In the one-loop approximation, the gaugino mass contributions are diagonal and the same for all three generations, so they can be absorbed into the common scalar masses and do not affect the diagonalization. The corrections to scalar masses at $M_G$ have the following leading flavor dependence $$ \eqalignno{ \Delta {\bf{m}}^2_E &\propto 2 {\fam=9{\mathchar"7115 } }_E{\fam=9{\mathchar"7115 } }_E^\dagger + 3 {\fam=9{\mathchar"7115 } }_{eu} {\fam=9{\mathchar"7115 } }^\dagger_{eu}\cr &= 5 \pmatrix{ \overline{z^2_e} C^2 & \overline{z_e y_e}CE^* & \overline{ z_ex_e}CB\cr \overline{z_ey_e} CE & \overline{z'^{2}_e}C^2 + \overline{y^2_e} \abs{E}^2 + \overline{x^2_e} B^2 & \overline{ y_ex'_e}EB + \bar{x}_e BA\cr \overline{z_ex'_e} CB & \overline{y_ex'_e} E^*B + \bar{x}_e BA & \overline{x'^2_e} B^2 + A^2},&(7.10)\cr} $$ where the overline represents the weighted average of the Clebsch factors, $\overline{z^2_e} = {1\over 5} (2z^2_e + 3z^2_{eu})$ and so on. Because $\Delta {\bf{m}}^2_E$ is hierarchical, assuming no big $x, y$ Clebsches (ADHRS models have some big $z$ Clebsches), the rotation matrix which diagonalizes it can be given approximately as $$ U_E(M_G) \simeq \pmatrix{ 1 & \bar{S}_{E_1} & \bar{S}_{E3}\cr -\bar{S}_{E_1} e^{-i\widetilde{\phi}} & e^{-i\widetilde{\phi}} & \bar{S}_{E3}\cr \bar{S}_{E_1}\bar{S}_{E_2}e^{-i\widetilde{\phi}} - \bar{S}_{E3} & -\bar{S}_{E2} e^{-i\widetilde{\phi}} & 1}, \eqno(7.11) $$ where $$ \eqalignno{ \bar{S}_{E_2} &= {\bar{x_e}B\over A},\; \bar{S}_{E_3} = { \overline{z_ex'_e}CB\over A^2}\cr \bar{S}_{E_1} &= { \overline{z_ey_e} C\abs{E}\over \overline{z'^{2}_e} C^2 + \overline{y^2_e} \abs{E}^2 + (\overline{x^2_e} - \overline{x}^2_e)B^2 },\; e^{-i\widetilde{\phi}} = {E\over \abs{E} }. \cr} $$ Similarly, for other scalar masses the leading flavor dependent corrections at $M_G$ are $$ \eqalignno{ \Delta {\bf{m}}^2_L &\propto {\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E + 3{\fam=9{\mathchar"7115 } }^\dagger_{q\ell}{\fam=9{\mathchar"7115 } }_{q\ell} + {\fam=9{\mathchar"7115 } }^\dagger_{n\ell}{\fam=9{\mathchar"7115 } }_{n\ell},\cr \Delta{\bf{m}}^2_Q &\propto {\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }_U^\dagger + {\fam=9{\mathchar"7115 } }_D{\fam=9{\mathchar"7115 } }^\dagger_D + 2{\fam=9{\mathchar"7115 } }_{qq}{\fam=9{\mathchar"7115 } }^\dagger_{qq} + {\fam=9{\mathchar"7115 } }_{q\ell}{\fam=9{\mathchar"7115 } }^\dagger_{q\ell},\cr \Delta {\bf{m}}^2_U &\propto 2 {\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_U + {\fam=9{\mathchar"7115 } }^\dagger_{eu}{\fam=9{\mathchar"7115 } }_{eu} + 2 {\fam=9{\mathchar"7115 } }_{ud}{\fam=9{\mathchar"7115 } }^\dagger_{ud},\cr \Delta{\bf{m}}^2_D &\propto 2{\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_D + {\fam=9{\mathchar"7115 } }^\dagger_{nd}{\fam=9{\mathchar"7115 } }_{nd} + 2{\fam=9{\mathchar"7115 } }^\dagger_{ud}{\fam=9{\mathchar"7115 } }_{ud},&(7.12) \cr} $$ and the rotation matrices which diagonalize them are given by expressions similar to (7.11) with Clebsches replaced by the appropriate ones. Then, the mixing matrices appearing at the lepton-slepton-gaugino vertices are given by $$ \eqalignno{ &W_{E_L}(M_G) = U^\dagger_L V_{E_L} \cr &\simeq \pmatrix{ 1 & S_{E_{L1}} - {S}_{L1} e^{i(\widetilde{\phi} - \phi_e)} & \bar{S}_{L_1} (\bar{S}_{L_2} -S_{E_{L_2}}) e^{i\widetilde{\phi}} - \bar{S}_{L_3}\cr (\bar{S}_{L_1}- \bar{S}_{E_{L_1}}) e^{-i(\widetilde{\phi}-\phi_e)} & e^{i(\widetilde{\phi}-\phi_e)} & -(\bar{S}_{L_2} - S_{E_{L_2}}) e^{i\widetilde{\phi} }\cr -S_{E_{L_1}}(\bar{S}_{L_2} - S_{E_{L_2}}) e^{-i{\phi}_e} + \bar{S}_{L_3} & (\bar{S}_{L_2}-S_{E_{L_2}}) e^{-i\phi_e} & 1 },\cr &&(7.13)\cr} $$ $$ \eqalignno{ &W_{E_L}(M_G) = U^\dagger_E V_{E_R} \cr &\simeq \pmatrix{ e^{i\phi_e} & S_{E_{R1}}e^{i{{\phi}}} e-\bar{S}_{E_{1}}e^{i\widetilde{\phi}}& \bar{S}_{E_1}(\bar{S}_{E_2} - S_{E_{R2}}) e^{i\widetilde{\phi}} - \bar{S}_{E_3}\cr \bar{S}_{E_1}e^{i\phi_e}-S_{E_{R1}}e^{i\widetilde{\phi}} & e^{i\widetilde{\phi}} & - (\bar{S}_{E_2}- S_{E_{R_2}})e^{-i{\phi}}\cr -{S}_{E_{R1}}(\bar{S}_{E_2} - S_{E_{R2}})+ \bar{S}_{E_3} e^{i{\phi_e}} & \bar{S}_{E_2} - S_{E_{R2}} &1},\cr &&(7.14)\cr} $$ \noindent where $$ \eqalignno{ S_{E_{L_1}} &= {z'_e C\over E'_e},\; S_{E_{R_1}} = {z_e C\over E'_e},\; E'_e = |y_e E - {x'_e x_e B^2\over A}|,\cr S_{E_{L_2}} &= {x'_e B\over A},\; S_{E_{R_2}} = {x_e B\over A},\; \phi_e = arg \left( y_e\abs{E} e^{i\widetilde{\phi}} - {x'_e x_e B^2\over A}\right)\cr & {\hskip 1.75in} \simeq \widetilde{\phi}\;\; (\hbox{if}\; y_e \sim x_e, x'_e),\cr \bar{S}_{E_1} &= { \overline{z_ey_e} C |E|\over \overline{z'^2_e} C^2 + \overline{y^2_e} \abs{E}^2 + (\overline{x^2_e} - \overline{x}^2_e) B^2}, \; \bar{S}_{E_2}={\bar{x}_eB\over A}, \; \bar{S}_{E_3} = {\overline{z_e x'_e} C B\over A^2},\cr \bar{S}_{L_1} &= { \widehat{z'_e y_e} C|E|\over \widehat{z^2_e} C^2 + \widehat{y^2_e} |E|^2+ (\widehat{x^2_e} - \widehat{x}^2_e)B^2 },\; \bar{S}_{L_2} = {\widehat{x'_e} B\over A},\; \bar{S}_{L_3} = { \widehat{z'_e x_e} CB\over A^2}, \cr \hbox{and}&\cr \bar{x}_e & = {1\over 5} (2x_e + 3 x_{eu}),\;\; \widehat{x'_e}= {1\over 5} (x'_e + 3 x'_{g\ell} + x'_{n\ell}) \;\; \hbox{etc.} \cr} $$ Note that $$ \bar{S}_{L_3} \sim \ {\hbox{Clebsch}}\ \times {CB\over A^2},\; S_{E_{L_1}} (\bar{S}_{L_2}-\bar{S}_{E_{L_2}}) = {\hbox{Clebsch}} \times {CB\over EA}.\eqno(7.15) $$ If there is no very big or small Clebsch involved and no accidental cancellation, $\bar{S}_{L_3}, \bar{S}_{E_3}$ can be neglected in $W$'s. Compared with $V_{KM}$, $$ \eqalignno{ &V_{KM} (M_G) = V^\dagger_{U_L} V_{D_L} \cr &\simeq \pmatrix{ 1 & S_{D_{L_1}}-S_{U_{L_1}} e^{-i(\phi_d-\phi_u)} & -S_{U_{L_1}} (S_{D_{L_2}}-S_{U_{L_2}}) e^{i\phi_u}\cr S_{U_{L_1}} - S_{D_{L_1}} e^{-i(\phi_d-\phi_u)} & e^{-i(\phi_d -\phi_u)} & (S_{D_{L_2}} - S_{U_{L_2}}) e^{i\phi_u}\cr S_{D_{L_1}} (S_{D_{L_1}} -S_{U_{L_2}}) e^{-i\phi_d} & -(S_{D_{L_2}} - S_{U_{L_2}}) e^{-i\phi_d} & 1 },\cr &&(7.16)\cr} $$ where $$ \eqalignno{ S_{U_{L1}} &= {z_u C\over E'_u},\; E'_u = \abs{ y_u E-{x_u x'_uB^2\over A} }, \phi_u = arg \left( y_u E- { x_u x'_u B^2\over A }\right),\cr S_{D_{L1}} &= {z_d C\over E'_d},\; E'_d = \abs{ y_d E - {x_dx'_d B^2\over A} },\; \phi_d = arg \left( y_dE - {x_dx'_d B^2\over A }\right),\cr S_{U_{L2}} &= {x_u B\over A},\cr S_{D_{L2}} &= {x_d B\over A},\cr} $$ we can see that the $W$'s and $V_{KM}$ do have similar hierarchical patterns, but have different Clebsch factors associated with their entries. When a specific model is given, one can calculate all the Clebsch factors and make some definite predictions for that particular model. For example, the ADHRS Model 6, which gives results in good agreement with the experimental data, has the following four effective fermion mass operators $$ \eqalignno{ O_{33} &= 16_3\; 10\; 16_3,\cr O_{23} &= 16_2\; {A_Y\over A_X}\; 10\; {A_Y\over A_X} 16_3,\cr O_{23} &= 16_2 {A_X\over M} \; 10\; {A_{B-L}\over A_X}\; 16_2 \;\;\; {\hbox{or other 5 choices}},\cr O_{12} &= 16_1 \left( {A_X\over M }\right)^3\; 10\; \left( {A_X \over M}\right)^3 16_2,&(7.17) \cr} $$ where $A_X, A_Y, A_{B-L}$ are adjoint's of $SO(10)$ with VEV's in the $SU(5)$ singlet, hypercharge, and $B-L$ directions. There are six choices of $O_{22}$ operators which give the same predictions for the fermion masses and mixings, but different Clebsches for other operators appearing above $M_G$. Fortunately, they do not enter the leading terms of the most important mixing matrix elements $W_{E_{L32}}, W_{E_{L31}}, W_{E_{R32}}, W_{E_{R31}}$, which appear in the leading contributions to the amplitudes of LFV processes and the electric dipole moment. The magnitude of the mixing matrix elements $V_{KM 32}, V_{KM 31}, W_{E_{L32}}, W_{E_{L31}},\\ W_{E_{R32}}, W_{E_{R31}}$, and the relevant Clebsch factors are listed in Tables 2 and 3. \begin{center} {\large\bf{Table 2}} \vskip .20in \begin{tabular}{|c|c|c|c|c|c|c|} \hline &$u$ & $d$ & $e$ &$eu$ & $q\ell$ &$ n\ell$\cr \hline $x$& $-1$ & $-{1\over 6}$ & ${3\over 2}$ & $-6$ & ${1\over 4}$ & 0\cr \hline $x'$ & $-1$ & $-{1\over 6}$ & ${3\over 2}$ & $-6$ & ${1\over 4}$ & 0\cr \hline $y$ & 0 & 1 & 3 & - & -& -\cr \hline $z$ & $-{1\over 27}$ & 1 & 1 & $-{ 1\over 27}$ & 1& 125\cr \hline $z'$&$-{1\over 27}$ & 1 & 1 & $-{1\over 27}$ & 1 & 125\cr \hline \multicolumn{7}{|c|}{$\widehat{x'_e} = {9\over 20},\; \overline{x_e} =- 3$}\cr \hline \end{tabular} \end{center} {\bf Table 2:} Clebsch factors for Yukawa coupling matrices in ADHRS \mbox{model 6}. \newpage \medskip \centerline{\bf{Table 3}} \vskip .20in \begin{center} \begin{tabular}{|c|c|c|c|} \hline &ADHRS models&Model 6&Relevant process\cr \hline $\abs{W_{E_{L_{32}}}/V_{ts} }$& $\abs{ {\widehat{x}'_e - x'_e\over x_d - x_u} }$ & 1.26&\cr &&&\cr \hline $\abs{W_{E_{R32}}/V_{ts} }$ & $\abs{ {\bar{x}_e -x_e\over x_d - x_u} }$ & 5.4 &\cr &&&\cr \hline $\abs{ W_{E_{L31}}/V_{td} }$ & $\abs{ {z'_e y_d (\widehat{x}'_e - x'_e)\over z_d y_e (x_d - x_u)} }$ & 0.42 &\cr &&&\cr \hline $\abs{ W_{E_{R31}}/V_{td} }$ & $\abs{ {z_e y_d (\bar{x}_e - x_e) \over z_d y_e (x_d - x_u)} }$& 1.8&\cr &&&\cr \hline $\abs{ { W_{E_{L32}}W_{E_{R31}}\over V_{ts} V_{td} } }$ & $\abs{ {z_e y_d(\bar{x}_e-x_e)(\widehat{x}'_e - x'_e) \over z_d y_e (x_d-x_u)^2 }}$ & 2.268 & $\mu\to e\gamma$ amplitude\cr &&&\cr \hline $\abs{ {W_{E_{R32}} W_{E_{L31}}\over V_{ts}V_{td} } }$ & $\abs{ {z'_e y_d (\bar{x}_e - x_e)(\widehat{x}'_e-x'_e)\over z_d y_e (x_d - x_u)^2 } } $& 2.268 &$\mu \to e\gamma$ amplitude\cr &&&\cr \hline $ \Bigg| \left( { \sqrt{2} W_{E_{L31}} W_{E_{R32}}\over \sqrt{ | W_{E_{L32}} W_{E_{R31}}|^2 + |W_{E_{R32}} W_{E_{L31}} |^2} } \right) / \left( { V_{td}\over V_{ts} }\right) \Bigg|$ & $ \bigg| { \sqrt{2} z'_e z_e y_d\over \sqrt{ (z^2_e + z'^2_e)} z_d y_e }\bigg| $ & ${1\over 3} $ & $d_e$\cr \hline \end{tabular} \end{center} {\bf Table 3:} The relevant Clebsch factors for $\mu \to e\gamma$ and $d_e$ in ADHRS \mbox{model 6}. \vskip .4in In ADHRS models $\tan\beta$ is large. The $\mu \to e\gamma$ rate for large $\tan\beta$ has been calculated in Sec.\ V and VI for $W_{E_{L32}} = W_{E_{R32}} = V_{ts}$ and $W_{E_{L31}} = W_{E_{R31}} = V_{td}$. To obtain the predictions of ADHRS models we only have to multiply the results by the suitable Clebsch factors. The relevant Clebsch factors for Model 6 are listed in Table 3. For a generic realistic GUT model with small tan$\beta$, for example the modified ADHRS models in which the down type Higgs lies predominantly in some fields which do not interact with the three low energy generations and contain only a small fraction of the doublets in the 10 which interact with the low energy generations \cite{CDW}, most of analysis should still hold. In this case the leading contributions to $\mu \to e\gamma$ are the same ones as in the minimal $SO(10)$ model of Ref. \cite{BHS} (Fig.\ 10 $b_{L,R}, c_{L,R}, c'_{L,R}$ of \cite{BHS}). The diagrams $c_{LR}, c'_{LR}$ involve the corrections to the trilinear scalar couplings. In the one-loop approximation the leading corrections to ${\fam=9{\mathchar"7110 } }_E$ at $M_G$ contain pieces proportional to ${\fam=9{\mathchar"7115 } }_E, {\fam=9{\mathchar"7115 } }_E ({\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E + 3 {\fam=9{\mathchar"7115 } }^\dagger_{q\ell} {\fam=9{\mathchar"7115 } }_{q\ell} + {\fam=9{\mathchar"7115 } }^\dagger_{n\ell} {\fam=9{\mathchar"7115 } }_{n\ell}), (2 {\fam=9{\mathchar"7115 } }_{E} {\fam=9{\mathchar"7115 } }^\dagger_{E}+ 3 {\fam=9{\mathchar"7115 } }_{eu} {\fam=9{\mathchar"7115 } }^\dagger_{eu}){\fam=9{\mathchar"7115 } }_E$ respectively. The piece proportional to ${\fam=9{\mathchar"7115 } }_E$ can be absorbed into ${\fam=9{\mathchar"7110 } }_{E_0}$ by a redefinition of $A_E$, the other two pieces are proportional to the product of ${\fam=9{\mathchar"7115 } }_E$ and the corrections to the scalar masses, $$ \eqalignno{ \Delta{\fam=9{\mathchar"7110 } }_E &= \Delta{\fam=9{\mathchar"7110 } }_{E_R} + \Delta{\fam=9{\mathchar"7110 } }_{E_L}\cr \Delta{\fam=9{\mathchar"7110 } }_{E_R} &= {1\over \mu_{E_R}} \Delta {\bf{m}}_E^2 \;{\fam=9{\mathchar"7115 } }_E\cr \Delta{\fam=9{\mathchar"7110 } }_{E_L} &= {1\over \mu_{E_L}} {\fam=9{\mathchar"7115 } }_E \;\Delta{\bf{m}}_{L}^2&(7.18)\cr} $$ where $\mu_{E_{L}}, \mu_{E_{R}}$ are proportional constants $(\mu_{E_R} = \mu_{E_L} = {6m^2_0 + A^2_0 \over 3A_0}$ in one-loop approximation). The LFV couplings in Fig.\ 2e, $\: \widetilde{e}^T_R U^T_E \Delta {\fam=9{\mathchar"7110 } }_E U_L \widetilde{e}_L v_D$, now can be written as $$ \eqalignno{ &{1\over \mu_{E_R}} \widetilde{e}^T_R U^T_E \Delta {\bf{m}}^2_E {\fam=9{\mathchar"7115 } }_E U_L \widetilde{e}_L v_D + {1\over \mu_{E_L}} \widetilde{e}^T_R U^T_E {\fam=9{\mathchar"7115 } }_E \Delta{\bf{m}}^2_L U_L \widetilde{e}_L v_D\cr = & {1\over \mu_{E_R}} \widetilde{e}_R^T \Delta{\overline{\bf{m}}}^2_E W^*_{E_R} \bar{\fam=9{\mathchar"7115 } }_E W^\dagger_{E_L} \widetilde{e}_L v_D + {1\over \mu_{E_L}} \widetilde{e}^T_R W^*_{E_R} \bar{\fam=9{\mathchar"7115 } }_E W^\dagger_{E_L} \Delta{\overline{\bf{m}}}^2_L \widetilde{e}_L v_D,&(7.19)\cr} $$ where the overline means that the matrix is diagonal. Again, the amplitudes are given by the same formulas as in \cite{BHS} (eqn.\ 29, 30), except that $V^e_{32}V^e_{31}(V^{e*}_{33})^2$ has to be replaced by $W_{E_{L32}} W_{E_{R31}} W^*_{E_{L33}} W^*_{E_{R33}}$, and $W_{E_{R32}} W_{E_{L31}}W^*_{E_{R33}}W^*_{E_{L33}}$, and ${5\over 7} I'_G$ by ${1\over \mu_{E_{R}}} {\Delta}\overline{m}^2_{E33}$ and ${1\over \mu_{E_L}} {\Delta}\overline{m}^2_{L33}$. The results in \cite{BHS} are only modified by some multiplicative factors and therefore represent the central values for the LFV processes. It was pointed out in \cite{BHS,DH} that the electric dipole moment of the electron $(d_e)$ constitutes an independent and equally important signature for the $SO(10)$ unified theory as $\mu \to e \gamma$ does. The diagrams which contribute to the electric dipole moment of the electron are the same as the ones which contribute to $\mu \to e\gamma$, with $\mu_L(\mu^c_L)$ replaced by $e_L(e_L^c)$. Thus a simple relation between $d_e$ and the $\mu\to e\gamma$ rate was obtained in the minimal $SO(10)$ model \cite{BHS}, $$ \Gamma (\mu \to e\gamma) = {\alpha\over 2} m^3_\mu |F_2|^2,\eqno(7.20) $$ $$ |d_e| = e|F_2| \abs{ {V_{td}\over V_{ts}}} \sin\phi = e \sqrt{ {2\Gamma (\mu\to e\gamma)\over \alpha m^3_\mu}}\abs{ {V_{td}\over V_{ts} } } \sin\phi,\eqno(7.21) $$ where $\phi$ is an unknown new CP violating phase defined by $$ Im [m_\tau(V^e_{31})^2 (V^{e*}_{33})^2] = |m_\tau(V^e_{31})^2 (V^{e*}_{33})^2| \sin \phi. $$ In a more generic $SO(10)$ model, such as the ADHRS model, we still have this simple relation but the mixing matrix elements have to be replaced by the $W$'s: $$ |d_e| = e \sqrt{ {2\Gamma (\mu \to e\gamma)\over \alpha m^3_\mu} } { \sqrt{2}|W_{E_{L31}}W_{E_{R31}}|\over \sqrt{ |W_{E_{L32}} W_{E_{R31}}|^2 + |W_{E_{R31}}W_{E_{L31}}|^2} } \sin\phi',\eqno(7.22) $$ where $\phi'$ is defined by $$ Im[m_\tau W_{E_{L31}}W_{E_{R31}} W^*_{E_{R33}}W^*_{E_{L33}}]= |m_\tau W_{E_{L31}}W_{E_{R31}}W^*_{E_{L33}}W^*_{E_{R33}}| \sin \phi'. $$ In particular, in ADHRS models there is only one CP violating phase, so the phase $\phi'$ can be related to the phase appeared in the KM matrix of the Standard Model. {}From eqn. (7.13) (7.14) (7.16) we can see that $\phi' \approx \phi_e, \phi_e \approx \phi_d \approx \widetilde{\phi}, \phi_u =0$ (because $y_u =0$). The rephrase invariant quantity $J$ of the KM matrix is given by $$ \eqalignno{ J &= Im V_{ud} V_{tb} V^*_{td} V^*_{ub}\cr &\simeq - S_{U_{L1}} S_{D_{L1}} (S_{D_{L2}} - S_{U_{L2}})^2\sin \phi_d.&(7.23)\cr} $$ Therefore the CP violating phase appeared in $d_e$ related to the CP violation in the Standard Model by $$ \sin \phi' \simeq {J\over |V_{td}| |V_{ub}|}.\eqno(7.24) $$ Finally, as mentioned in the Sec.\ III, we consider the possibility that the slight non-degeneracy between the first two generation scalar masses could give a significant contribution to the flavor changing processes because of the larger mixing matrix elements. We still use ADHRS models as an example to estimate this contribution to the LFV process $\mu \to e\gamma$. For an order of magnitude estimate, the mass insertion approximation in the super-KM basis employed in \cite{BH} will serve as a convenient method. After rotating the $\Delta {\bf{m}}^2_E$ in eqn.\ (7.10) to the charged lepton mass eigenstate basis, the contribution from the first two generations to ${\Delta m^2_{E_{21}}}$ is $$ \eqalignno{ {\Delta}{m}^2_{E_{21}} (2{\hbox{-}}1) \simeq V_{E_{R22}} V^*_{E_{R21}} \Delta m^2_{E22} + V_{E_{R12}} V^*_{E_{R11}} \Delta m^2_{E_{11}} + V_{E_{R22}} V^*_{E_{R11}} \Delta m^2_{E_{21}}\cr \simeq [- {z_e C\over E'_e} { \overline{z'^2_e} C^2 + \overline{{y}^2_e} |E|^2 + \overline{x^2_e} B^2\over A^2} + {z_eC\over E'_e} {\overline{z^2_e} C^2\over A^2} + e^{-i\phi_e} {\overline{z_e {y}_e}} {CE\over A^2} ] \Delta m^2_{E_{33}}\cr \simeq - {z_e\over y_e} \left[ {\overline{y^2_e} C|E| + \overline{x^2_e} {CB^2\over |E|} + \overline{z_ey_e}C|E|\over A^2} \right] \Delta m^2_{E_{33}} \hspace{.7in}\cr {(\hbox{assume }} \; z_e = z'_e \; {\hbox{as in ADHRS model}}). &(7.25) \cr} $$ Compared with the result found in \cite{BH} for minimal $SU(5)$: $$ \eqalignno{ \Delta m^2_{E_{21}} (BH) &= V^*_{ts} V_{td} \Delta m^2_{E_{33}} \cr &\simeq - {z_d C\over E'_d} {(x_d - x_u)^2B^2\over A^2} \Delta m^2_{E_{33}}\cr &\simeq - {z_d\over y_d} {(x_d - x_u)^2 {CB^2\over |E|}\over A^2} \Delta m^2_{E_{33}}, &(7.26) \cr} $$ we can see that if the Clebsch factors are $O(1)$, this contribution is comparable to that of the minimal $SU(5)$ model. In order for this contribution to be competitive with the dominant diagrams (Fig.\ 10 $b_{L,R}, c_{L,R}, c'_{L,R} $ of \cite{BHS}) which are enhanced by ${m_\tau\over m_\mu}$, large Clebsch factors are required. While it is possible to have large Clebsch factors, we consider them as model dependent, not generic to all realistic unified theories. \newpage \noindent{\bf VIII. Conclusions} In supersymmetric theories, the Yukawa interactions which violate flavor symmetries not only generate the quark and lepton mass matrices, but necessarily also lead to radiative breaking of flavor symmetries in the squark and slepton mass matrices, leading to a variety of flavor signals. While such effects have been well studied in the MSSM and, more recently, in minimal unified models, the purpose of this paper has been to explore these phenomena in a wide class of grand unified models which have realistic fermion masses. We have argued that, if the hardness scale $\Lambda_H$ is above $M_G$, the expectation for all realistic grand unified supersymmetric models is that non-trivial flavor mixing matrices should occur at {\it all} neutral gaugino vertices. These additional, weak scale, flavor violations are expected to have a form similar to the Kobayashi-Maskawa matrix. However, the precise values of the matrix elements are model dependent and have renormalization group scalings which differ from those of the Kobayashi-Maskawa matrix elements. It is the non-triviality of the flavor mixing matrices of neutral gaugino couplings in the up quark sector which strongly distinguishes between the general and minimal unified models, as shown in Table 1. Although the minimal unified models provide a simple approximation to flavor physics, they are not realistic, so we stress the important new result that flavor mixing in the up sector couplings of neutral gauginos is a necessity in unified models. this leads to four important phenomenological consequences. While the $D^0- \bar{D}^0$ mixing induced by this new flavor mixing is generally not close to the present experimental limit, it could be much larger than that predicted in the standard model. The new mixing in the up-quark sector implies that there may be significant radiative contributions to the up quark mass matrix which arise when the superpartners are integrated out of the theory. This is illustrated in Figure 4, where the new mixing matrix elements have been taken to be a factor of three larger than the corresponding Kobayashi-Maskawa matrix elements. In this case the entire up quark mass could be generated by such a radiative mechanism: above the weak scale the violation of up quark flavor symmetries lies in the squark mass matrix. The electric dipole moment of the neutron, $d_n$, is a powerful probe of the neutral gaugino flavor mixing induced by unified theories. In the minimal $SO(10)$ theory, $d_n$ arises from the flavor mixing in the down sector, which leads to a down quark dipole moment, $d_d$. However, in realistic models the flavor mixing in the up quark sector leads to a $d_u$ which typically provides the dominant contribution to $d_n$. Thus the neutron electric dipole moment is a more powerful probe of unified supersymmetric theories than previously realized. The presence of flavor mixing in the up sector plays a very important role in determining the branching ratio for a proton to decay to $K^0 \mu^+$. In the minimal models, without such mixings, this branching ratio is expected to be about $10^{-3}$: the charged lepton mode will not be seen and experimental efforts must concentrate on the mode containing a neutrino, $K^+ \nu$. However, including these mixings the charged lepton branching ratio is greatly increased to about $0.1$. While this number is very model dependent, we nevertheless think that this effect greatly changes the importance of searching for the charged lepton mode. These four phenomenologocal consequences are sufficiently interesting that we stress once more that they appear as a necessity in a wide class of unified theories. The absence of mixing in the up sector is a special feature of the minimal models. Since the flavor sectors of the minimal models must be augmented to obtain realistic fermion masses, any conclusions based on the absence of flavor mixings in the up sector are specious. A second topic addressed in this paper is the effect of large $\tan \beta$ on the lepton process, $\mu \rightarrow e \gamma$ which is expected in unified supersymmetric $SO(10)$ models. The amplitude for this process has a contribution proportional to $\tan \beta$. In this paper, we have found that the naive expectation that large $\tan \beta$ in supersymmetric $SO(10)$ is excluded by $\mu \rightarrow e \gamma$ is incorrect, at least for all values of the superpartner masses of interest. Contour plots for the $\mu \rightarrow e \gamma$ branching ratio are shown in Figures 7 and 8. It depends sensitively on the parameter $\Delta$, which is the mass splitting between the scalar electron and scalar tau, and is plotted in Figure 9. Lower values of the top quark Yukawa coupling, which for large $\tan \beta$ still give allowed predictions for the $b/\tau$ mass ratio, give a much reduced value for $\Delta$, thereby reducing the $\mu \rightarrow e \gamma$ rate and partially compensating the $\tan^2 \beta$ enhancement. A further significant suppression of an order of magnitude is induced by the renormalization group scaling of the leptonic flavor mixing angles, and is shown in Figure 10. The net effect is that while the case of $\tan \beta \approx m_t/m_b$ is not excluded in SO(10), the $\mu \rightarrow e \gamma$ rate is still typically larger than for moderate $\tan \beta$, so that this process provides a more powerful probe of the theory as $\tan \beta $ increases. For large $\tan \beta$, $\mu$ and $M_2$ become the physical masses of the two charginos. The $\mu \rightarrow e \gamma$ contours of Figure 8 show that $\mu$ and $M_2$ should not be too large, providing an important limit to the chargino masses in the large $\tan \beta$ limit. Furthermore, this constrains the LSP mass to be quite small. We find that in this region it is still possible for the LSP to account for the observed dark matter, and even to critically close the universe, as can be seen from Figure 11. However, the requirement that the LSP mass be larger than 45 GeV suggests that the two light charginos will not be light enough to be discovered at LEP II. As an example of theories with both a realistic flavor sector and large $\tan \beta$ we studied the models introduced by Anderson et al. The flavor sectors of these theories are economical: the free parameters can all be fixed from the known quark and lepton masses and mixings. Hence the flavor mixing matrices at all neutral gaugino vertices can be calculated. These are shown for the lepton sector of model 6 in Table 3. The Clebsch factors enhance the $\mu \rightarrow e \gamma$ amplitude by a factor of 2.3, and suppress $d_e$ by a factor of 3. Even taking the top quark Yukawa coupling to have its lowest value the rate for $\mu \rightarrow e \gamma$ in this theory is very large. Another interesting feature of these theories is that the flavor sectors contain just a single CP violating phase. This means that the phase which appears in the result for $d_n$ and $d_e$ can be computed: since it is closely related to the phase of the Kobayashi-Maskawa matrix it is not very small. That which appears in $d_e$ is given in eqn.\ (7.24) and is numerically about 0.2. We have computed the radiative corrections to $m_u$ in the ADHRS models and have found that the new mixing matrices in the up sector are not large enough to yield sizable contributions: thus the ADHRS analysis of the quark mass matrices is not modified. Furthermore, due to a cancellation special to these theories, there is no contribution to $d_n$ from the up quark at one loop. \newpage \noindent {\bf Acknowledgements} The authors would like to thank Hitoshi Murayama for many useful discussions. This work was supported in part by the Director, Office of Energy Research, Office of High Energy and Nuclear Physics, Division of High Energy Physics of the U.S. Department of Energy under contract No. DE-AC03-76SF00098 and in part by the National Science Foundation under Grant No. PHY-90-21139. The work of N.A-H was supported by an NSERC '67 fellowship. \vskip .6in \noindent Note Added: While finalizing this work, we received a preprint by Ciafaloni, Romanio and Strumia\cite{CRS}, where the large $\tan\beta$ scenario is also considered. However, unlike this work, they assume strict universality in soft scalar masses, such that imposing electroweak symmetry breaking leads them into a region of parameter space with a high mass (1 TeV) for the sleptons. In their discussion of general models, they do not include flavor violating RG scaling of scalar masses above $M_G$. \newpage \noindent{\bf{Appendix}} In this appendix, we first give a more complete treatment of mixing matrix scaling in the lepton sector, and then give a treatment for the quark sector. Let us return to (5.7) and consider the effect of including the $({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i}$ term. In general the scaling from $M_{PL}$ to $M_G$ will generate a ${\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E$ not diagonal in the same basis as ${\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E$, so we expect some non-zero $({\fam=9{\mathchar"7110 } }^\dagger_{E}{\fam=9{\mathchar"7110 } }_E)_{3i}$. {}From the RGE for ${\fam=9{\mathchar"7110 } }_E$, neglecting gauge couplings, $$ - {d\over dt} {\fam=9{\mathchar"7110 } }_E = {\fam=9{\mathchar"7110 } }_E [ 5{\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E + Tr (3{\fam=9{\mathchar"7115 } }^\dagger_D {\fam=9{\mathchar"7115 } }_D + {\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E)] + {\fam=9{\mathchar"7115 } }_E [4{\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E + Tr (6{\fam=9{\mathchar"7110 } }_D {\fam=9{\mathchar"7115 } }_D^\dagger + 2 {\fam=9{\mathchar"7110 } }_E{\fam=9{\mathchar"7115 } }_E^ \dagger)].\eqno(A.1) $$ We have $$ \eqalignno{ -{d\over dt} ({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E) &=5[{\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E {\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E + {\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E{\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E]+ 2Tr(3{\fam=9{\mathchar"7115 } }^\dagger_D {\fam=9{\mathchar"7115 } }_D + {\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E) {\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E\cr &+ 8{\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E{\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E + ({\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E + {\fam=9{\mathchar"7115 } }_E^\dagger{\fam=9{\mathchar"7110 } }_E) Tr(6{\fam=9{\mathchar"7110 } }_D{\fam=9{\mathchar"7115 } }^\dagger_D + 2{\fam=9{\mathchar"7110 } }_E{\fam=9{\mathchar"7115 } }^\dagger_E).&(A.2) \cr} $$ Then, to first order in the off diagonal parts of ${\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E$ and ${\fam=9{\mathchar"7110 } }_E{\fam=9{\mathchar"7110 } }_E^\dagger$, and keeping only third generation Yukawa couplings we have $$ - {d\over dt} ({\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E)_{3i} = ({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i}[17 {\lambda}^2_\tau + 6 {\lambda}^2_b + 6\eta {\lambda}_b {\lambda}_\tau],\eqno(A.3) $$ where $\eta \equiv {\zeta_{D_{33}}\over \zeta_{E_{33}} }$. Because of the large numerical coefficient in front of $\lambda^2_\tau, \lambda^2_b$ in the above equation, (${\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i}$ is driven to zero more rapidly than $W_{L3i}$, after which it ceases to have any effect on the running of $W_{L3i}$. More explicitly, from (5.7) we have that $$ {d\over dt} (m^2_{L3i}(t)e^{\int^t_0 dt'\lambda^2_\tau (t')}) =- 2 ({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i}(t) e^{\int^t_0 dt' \lambda^2_\tau(t')}. \eqno(A.4) $$ Solving (A.3) for $({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i}(t)$ and inserting into (A.4) we get $$ -{d\over dt} \left( m^2_{L 3i}(t)e^{\int^t_0 dt'\lambda^2_\tau (t')}\right) = 2({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i}(M_G)e^{-\int^t_0 dt'[16 \lambda^2_\tau +6\lambda^2_b + 6\eta \lambda_b \lambda_t](t')}.\eqno(A.5) $$ Integrating (A.5), we find $$ \eqalignno{ m^2_{L 3i} (M_S)e^{I_\tau} - m_{L 3i}^2 (M_G) &= -2 \int_0^{ {1\over 16\pi^2}\log {M_G\over M_S}} dt e^{-\int^t_0 dt'[16\lambda^2_\tau + 6 \lambda^2_b + 6\eta \lambda_b\lambda_\tau](t')}\cr &\times ({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i}(M_G)\cr &\equiv \delta ({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i} (M_G).&(A.6)\cr} $$ So, we have $$ m^2_{L 3i}(M_S) = e^{-I_\tau}[m^2_{L 3i} (M_G) + \delta ({\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E) (M_G)].\eqno(A.7) $$ We expect $m^2_{L 3i}$ and ${({\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_{E})}_{3i}$ to be related by some combination of Clebsches $x$ at $M_G$ as follows: $$ ({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i} = {A^2_0\over m^2_0} xm^2_{L 3i}\eqno(A.8) $$ Where $A_0, m^2_0$ are the universal $A$ parameter and scalar mass at $M_{PL}$, respectively. Then, we have from (A.7) $$ W^\dagger_{L33} W_{L 3i} (M_S) = e^{-I_\tau}{\Delta m^2(M_G)\over \Delta m^2 (M_S)} [1 + \delta {A^2_0\over m^2_0} x] W^\dagger_{L33}W_{L 3i} (M_G).\eqno(A.9) $$ Clearly if $\delta {A^2_0\over m^2_0} x$ $\ll 1$, inclusion of the $({\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E)_{3i}$ term in (5.7) do not change any of our results. If $\delta{A^2_0\over m^2_0} x$ $\sim 1 $ or $\gg 1$, we can still of course use (A.9), but the suppression effect may disappear. A simple estimate shows, however, that $\delta$ itself is already small $\sim {1\over 10}$, and so we are only in trouble if ${A^2_0\over m^2_0} x$ is big. To see this, replace $\lambda_\tau, \lambda_b$ and $\eta$ by some average values $\bar{\lambda}_\tau, \bar{\lambda}_b$ and $\bar{\eta}$ in the expression (A.6) for $\delta$. Then, $$ \eqalignno{ \delta &= -2 \int_0^{ {1\over 16 \pi^2}\log {M_G\over M_S}} e^{-t(16\bar{\lambda}_\tau^2 + 6 \bar{\lambda}^2_b + 6\bar{\lambda}_b \bar{\lambda}_t \bar{\eta} )}\cr &= - {1\over 8\bar{\lambda}^2_t + 3(\bar{\lambda}^2_b +\bar{\eta}\bar{\lambda}_b\bar{\lambda}_\tau)} \left[ e^{- {1\over 16\pi^2} \log {M_G\over M_S} (16 \bar{\lambda}^2_\tau + 6 \bar{\lambda}^2_b + 6\bar{\eta}\bar{\lambda}_b\bar{\lambda}_\tau)} \right]. &(A.10) \cr} $$ So, $$ |\delta | < { 1\over 8 \bar{\lambda}^2_\tau + 3(\bar{\lambda}^2_b +\bar{\eta} \bar{\lambda}_b \bar{\lambda}_\tau). }\eqno(A.11) $$ For the $\bar{\lambda}$'s between 0.5 and 1, and $\bar{\eta} \sim 1, |\delta |$ ranges from ${1\over 3}$ to ${1\over 15}$. How can we qualitatively understand the above results for the scaling of mixing matrices? The renormalization group equations try to align the soft supersymmetry breaking flavor matrices with whatever combination of flavor matrices responsible for their renormalization. However, because a given coupling can only be renormalized by harder couplings, there is a hierarchy in which flavor matrices affect the running of others. The Yukawa matrices, being dimensionless, can only be affected by other Yukawa matrices. In the lepton sector, this is the reason that the basis in which e.g. ${\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E$ is diagonal does not change. Next, the soft trilinear terms, having mass dimension one, can only be affected by other trilinear terms and Yukawa couplings. Again in the lepton sector this means that e.g. ${\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E$ tries to align itself with ${\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E$. Finally, the scalar mass, having dimension two, are affected by everything: ${\bf{m}}^2_L$ tries to align with ${\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E$, but suffers interference from ${\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E$, unless ${\fam=9{\mathchar"7110 } }^\dagger_E {\fam=9{\mathchar"7110 } }_E$ is diagonal in the same basis as ${\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E$. Even if ${\fam=9{\mathchar"7110 } }^\dagger_E{\fam=9{\mathchar"7110 } }_E$ is not diagonal in the same basis as ${\fam=9{\mathchar"7115 } }^\dagger_E {\fam=9{\mathchar"7115 } }_E$, it is trying to align itself with ${\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E$, so ${\bf{m}}^2_L$ will still tend to align with ${\fam=9{\mathchar"7115 } }^\dagger_E{\fam=9{\mathchar"7115 } }_E$. {}From the above discussion, it is clear that the situation is slightly complicated in the quark sector. In the lepton sector, there was a fixed direction in flavor space given by ${\fam=9{\mathchar"7115 } }_E$, with which the soft matrices aligned. In the quark sector, we have both ${\fam=9{\mathchar"7115 } }_U$ and ${\fam=9{\mathchar"7115 } }_D$, and ${\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }_U^\dagger$, ${\fam=9{\mathchar"7115 } }_D {\fam=9{\mathchar"7115 } }_D^\dagger$ are misaligned $(V_{KM}\neq 1)$. This complicates the analysis for $W_{U_L}, W_{D_L}$ so we discuss them last. Let us now examine the scaling of $W_{U_R}, W_{D_R}$. (Throughout the following, we assume degeneracy between first two generation scalar masses, we neglect all Yukawa coupling matrix eigenvalues except those of the third generation, and we do not include the effect of trilinear soft terms in the scaling. The last assumption is made for simplicity; we can make similar arguments about the importance of these neglected trilinear terms as we did above in the lepton sector.) First, we show that the basis in which ${\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_U$ is diagonal remains fixed. The RGE for ${\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_U$ is $$ - {d\over dt}{\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_U = 6({\fam=9{\mathchar"7115 } }^\dagger_U {\fam=9{\mathchar"7115 } }_U)^2 + 2{\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_D {\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_U + 2(3Tr{\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U - {16\over 3} g^2_3 -3 g^2_2 - {13\over 15} g^2_1) {\fam=9{\mathchar"7115 } }_U^\dagger {\fam=9{\mathchar"7115 } }_U.\eqno(A.12) $$ Working in a basis where ${\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_U$ is diagonal, let us see if ${d\over dt} {\fam=9{\mathchar"7115 } }^\dagger_U{\fam=9{\mathchar"7115 } }_U$ has off-diagonal components. We have, (recalling that in this basis ${\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_D = V_{KM}\bar{{\fam=9{\mathchar"7115 } }}^2_D V^\dagger_{KM})$, $$ \eqalignno{ - {d\over dt} ({\fam=9{\mathchar"7115 } }^\dagger_U {\fam=9{\mathchar"7115 } }_U)_{\stackrel{ij}{i\neq j}} &= 2 (\bar{{\fam=9{\mathchar"7115 } }}_UV_{KM} \bar{{\fam=9{\mathchar"7115 } }}^2_DV^\dagger_{KM} \bar{{\fam=9{\mathchar"7115 } }}_U)_{ij}\cr &= 2\bar{{\fam=9{\mathchar"7115 } }}_{U_i} V_{KM_{i\ell}} \bar{{\fam=9{\mathchar"7115 } }}_{D_{\ell}}^2 V^\dagger_{KM_{\ell j}} \bar{{\fam=9{\mathchar"7115 } }}_{U_j}\cr &= 0 \ \hbox{for} \ i, j\neq 3 &(A.13)\cr} $$ since we neglect all Yukawa's except the third generation. Similarly, the basis in which ${\fam=9{\mathchar"7115 } }^\dagger_D{\fam=9{\mathchar"7115 } }_D$ is diagonal does not change. Thus, the discussion for the scaling of $W_{U_R}, W_{D_R}$ is completely analogous to that in the lepton sector, and we find $$ W_{U_{R 3i}}W^\dagger_{U_{R 33}}(M_S) = e^{-2I_t} {\Delta m^2_U(M_G)\over \Delta m^2_U(M_S)} W_{U_{R 3i}}W^\dagger_{U_{R 33}}(M_G),\eqno(A.14) $$ $$ W_{D_{R 3i}}W^\dagger_{D_{R 33}}(M_S) = e^{-2I_b} {\Delta m^2_D(M_G)\over \Delta m^2_D(M_S)} W_{D_{R 3i}}W^\dagger_{D_{R 3}}(M_G).\eqno(A.15) $$ We now turn to $W_{U_L}, W_{E_L}$. Let $V^*_{U_L}(t)$ be the matrix diagonalizing ${\fam=9{\mathchar"7115 } }_U {\fam=9{\mathchar"7115 } }^\dagger_U(t)$: $$ {\fam=9{\mathchar"7115 } }_U {\fam=9{\mathchar"7115 } }^\dagger_U (t)= V^{*}_{U_{L}}(t) {\bar{{\fam=9{\mathchar"7115 } }}}^2_U (t) V^{*\dagger}_{U_{L}} (t).\eqno(A.16) $$ In the superfield basis in which ${\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U$ is diagonal, the squark mass matrix is $\widetilde{\bf{m}}^{2*}_{Q 3i} = V^\dagger_{U_L}{\bf{m}}^{2*}_Q V_{U_L}$. Note as before that $\widetilde{\bf{m}}^{2*}_{Q 3i} = (W^\dagger_{U_L}\overline{\bf{m}}^{2*}_Q W_{U_L})_{3i}= W_{U_{L 3i}}W^\dagger_{U_{L 33}}\Delta m^2_Q$, so we are interested in ${d\over dt} \widetilde{\bf{m}}^{2*}_{Q 3i}$. Now, $$ \eqalignno{ {d\over dt} \widetilde{\bf{m}}_Q^{2*} = {d\over dt}(V^\dagger_{U_L} {\bf{m}}^{2*}_QV_{U_L})&= \left( {d\over dt} V^\dagger_{U_L}\right) {\bf{m}}^{2*}_Q V_{U_L} + V^\dagger_{U_L} {d\over dt} {\bf{m}}^{2*}_Q V_{U_L} + V^\dagger_{U_L} {\bf{m}}^{2*}_Q {d\over dt} V_{U_L}\cr &= \left[ \widetilde{\bf{m}}^{2*}_Q, V^\dagger_{U_L} {d\over dt} V_{U_L}\right] +V^\dagger_{U_L} {d\over dt} {\bf{m}}^{2*}_Q V_{U_L}.&(A.17)\cr} $$ The second term is the analogue of what we have already seen in the lepton and right-handed quark sector; using the RGE for ${\bf{m}}^{2*}_Q$ we find to leading order $$ \left( V^\dagger_{U_L} {d\over dt} {\bf{m}}^{2*}_Q V_{U_L}\right)_{3i} =- (\lambda^2_t + \lambda^2_b) \widetilde{\bf{m}}^2_{Q 3i}.\eqno(A.18) $$ Now, $V^\dagger_{U_L} {d\over dt} V_{U_L}$ is obtained from the RGE for ${\fam=9{\mathchar"7115 } }_U {\fam=9{\mathchar"7115 } }^\dagger_U$. Actually, note that $$ V^\dagger_{U_L} \left( {d\over dt} {\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U\right) V_{U_L} = \left[ V^\dagger_{U_L} {d\over dt} V_{U_L}, \bar{{\fam=9{\mathchar"7115 } }}^2_U\right] + {d\over dt} \bar{{\fam=9{\mathchar"7115 } }}^2_U,\eqno(A.19) $$ so that only $[V^\dagger_{U_L} {d\over dt} V_{U_L}, \bar{{\fam=9{\mathchar"7115 } }}^2_U]$ is determined. (This is a reflection of the fact that $V_{U_L}$ is not unique: let $X (t)$ be any unitary transformation leaving $\overline{\bf{m}}^2_Q(t)$ invariant: $ \overline{\bf{m}}^2_Q(t) = X^\dagger (t) \overline{\bf{m}}^2_Q(t) X(t). $ In our case, $X(t)$ is most generally a $U(2)$ matrix in the first two generation subspace. Then, if $V_{U_L}$ diagonalizes ${\bf{m}}^{2*}_Q$, so does $V_{U_L}X$. Under this change, $V^\dagger_{U_L} {d\over dt} V_{U_L}$ is not invariant, but $[ V^\dagger_{U_L} {d\over dt} V_{U_L}, \bar{{\fam=9{\mathchar"7115 } }}^2_U]$ {\it {is}} invariant). Further, since we neglect first two generations Yukawa eigenvalues, $[V^\dagger_{U_L} {d\over dt} V_{U_L}, \bar{{\fam=9{\mathchar"7115 } }}^2_U]_{ij} =0 $ for $ i, j= 1,2$, and only $[V^\dagger_{U_L} {d\over dt} V_{U_L}, \bar{{\fam=9{\mathchar"7115 } }}^2_U]_{3i (i3)} = (\mp) \lambda^2_t V^\dagger_{U_L} {d\over dt} V^\dagger_{U_{L i 3 (3 i)}}$ is determined, and we can choose all other components of $V^\dagger_{U_L} {d\over dt} V_{U_L}$ to vanish. {}From the RGE for ${\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }_U^\dagger, $ $$ \eqalignno{ - {d\over dt} ({\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }_U^\dagger) &= 6({\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U)^2 + 2(3 Tr {\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }_U^\dagger - {16\over 3} g^2_3 - 3g^2_2 - { 13\over 15} g^2_1) {\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }_U^\dagger\cr &+ \{ {\fam=9{\mathchar"7115 } }_U {\fam=9{\mathchar"7115 } }^\dagger_U, {\fam=9{\mathchar"7115 } }_D {\fam=9{\mathchar"7115 } }^\dagger_D\},&(A.20)\cr} $$ we find $$ \eqalignno{ -\left(V^\dagger_{U_L} ({d\over dt}{\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_{U}) V_{U_L}\right)_{3i} &= \{ \bar{{\fam=9{\mathchar"7115 } }}_U^2, V_{KM} \bar{{\fam=9{\mathchar"7115 } }}^2_D V^\dagger_{KM}\}_{3i}\cr &= {\lambda}^2_t{\lambda}^2_bV_{KM 33}V^\dagger_{KM 3i}, &(A.21)\cr} $$ and thus $$ \left( V^\dagger_{U_L} {d\over dt} V_{U_{L 3i}}\right) =- \lambda^2_b V^\dagger_{KM 3i} V_{KM 33}.\eqno(A.22) $$ Thus to leading order $$ [V^\dagger_{U_L} {\bf{m}}^2_Q V_{U_L}, V^\dagger_{U_L} {d\over dt} V_{U_L}]_{3i} =- \Delta {{m}}^2_Q \,\lambda^2_b \,V^\dagger_{KM 3i} V_{KM 33},\eqno(A.23) $$ and finally we have $$ - {d\over dt} (W_{U_{L3i}} W^\dagger_{U_{L 33}} \Delta m_Q^2) = (\lambda^2_t + \lambda_b^2) W_{U_{L 3i}}W^\dagger_{U_{L 33}} \Delta {{m}}^2_Q + \lambda^2_b V^\dagger_{KM 3i} V_{KM 33}\Delta {{m}}^2_Q.\eqno(A.24) $$ Similarly we find $$ - {d\over dt} (W_{D_{L 3i}} W^\dagger_{D_{L 33}} \Delta {{m}}^2_Q) = (\lambda^2_t + \lambda^2_b) W_{D_{L 3i}} W^\dagger_{D_{L 33}} \Delta {{m}}^2_Q + \lambda^2_t V^\dagger_{KM 3i}V_{KM 33}\Delta {{m}}^2_{Q}.\eqno(A.25) $$ We can formally solve the above equations, e.g. $$ W_{U_{L 3i}} W^\dagger_{U_{L 33}} (M_S) = e^{-\left( I_t + I_b + \int^{M_G}_{M_S} dt' \lambda^2_b {V^\dagger_{KM 3i} V_{KM_{ 33}}\over W_{U_{L 3i}}W^\dagger_{U_{L 33}}} \right)} {\Delta {{m}}^2_Q (M_G)\over \Delta {{m}}^2_Q(M_S)} W_{U_{L 3i}} W^\dagger_{U_{L 33}} (M_G),\eqno(A.26) $$ and, to a good approximation, given that $W_{U_{L 3i}}$ does not scale very significantly, we can replace $$ \int^{M_G}_{M_S} dt' \lambda^2_b {V^{\dagger}_{KM 3i} V_{KM 33}\over W_{U_{L 3i}}W^\dagger_{U_{L 33}} } \approx I_b {V^\dagger_{KM 3i} V_{KM 33}\over W_{U_{L 3i}} W^\dagger_{U_{L 33}} } (M_G).\eqno(A.27) $$ So, an approximate solution of the RGE for $W_{U_L}, W_{D_L}$ is $$ W_{U_{L 3i}} W^\dagger_{U_{L 33}} (M_S) \approx e^{ -\left( I_t + I_b \left( 1+ { V^\dagger_{KM 3i} V_{KM_{ 33}}\over W_{U_{L 3i}} W^\dagger_{U_{L 33}}} (M_G)\right) \right) } {\Delta {{m}}^2_Q (M_G)\over \Delta {{m}}^2_Q(M_G)} W_{D_{L 3i}} W^\dagger_{U_{L 33}} (M_G),\eqno(A.28) $$ and similarly $$ W_{D_{L 3i}} W^\dagger_{D_{L 33}}(M_S) \approx e^{ -\left( I_b + I_t \left( 1+ { V^\dagger_{KM 3i} V_{KM 33} \over W_{U_{L 3i}} W^\dagger_{U_{L 33}}} (M_G) \right) \right) } {\Delta {{m}}^2_Q(M_G)\over \Delta {{m}}^2_Q(M_S)} W_{D_{L 3i}} W^\dagger_{D_{L 33}} (M_G).\eqno(A.29) $$ The above results are in agreement with qualitative expectations; the extra terms in the exponential of (A.28) and (A.29) are a reflection of the fact that the bases in which ${\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U$ and ${\fam=9{\mathchar"7115 } }_D{\fam=9{\mathchar"7115 } }^\dagger_D$ are diagonal change with scale. For moderate $\tan\beta$, however, we expect that the basis in which ${\fam=9{\mathchar"7115 } }_U{\fam=9{\mathchar"7115 } }^\dagger_U$ is diagonal should not change with scale, and in this limit the extra term drops out of (A.28). \newpage
\section{Experimental Methods} The films used in this study were grown using ion-beam sputtering (IBS) in a chamber with a base pressure of 2$\times$10$^{-8}$ torr. The deposition system is described in more detail elsewhere.\cite{fesiprb} All samples used in this study were grown at a substrate temperature of 200$^{\circ}$. All comparisons between films grown on glass and single-crystal substrates will be made on samples which were deposited {\em simultaneously} so as to eliminate any reproducibility issues. The substrates used in this study were glass coverslips, MgO (001), Ge(001) and Al$_2$O$_3$(11\={2}0). The MgO and Al$_2$O$_3$ substrates were cleaned according to a recipe reported by Farrow and coworkers.\cite{farrow} The glass and Ge substrates were rinsed in solvents. All films are capped with a 200\AA\ Ge oxidation barrier. The magnetic and structural properties of the films are stable for at least one year. \section{Structural Characterization} Figure~\ref{epihigh} shows high-angle x-ray diffraction spectra for a purely (001)-oriented (Fe40\AA/Si14\AA)x60 multilayer grown on MgO(001) and a purely (011)-oriented multilayer grown simultaneously on glass. While multiple superlattice satellites are observed around the Fe(002) peak for the film grown on MgO, the film on glass has only one peak corresponding to (011)-textured growth. A single superlattice satellite is typically observed in multilayers on glass on the low-angle side in keeping with previous observations.\cite{foiles2} An estimate of the crystallite sizes in these films can be derived using the Scherrer formula. This analysis gives a coherence length of 165\AA\ for the film on glass and 188\AA\ for the film on MgO. Since the crystalline coherence lengths of these films are similar, the presence of high-angle satellites in the film on MgO must be due to better layering. The small-angle x-ray scattering data shown in Figure~\ref{epilow} confirm this hypothesis. The multilayer on MgO has 4 low-angle peaks, indicating a moderate degree of composition modulation. The multilayer on glass shows only two relatively broad peaks, indicating larger interfacial roughness and less order in the layering. The low-angle x-ray spectra are consistent with rocking curves which are only about 1$^{\circ}$ wide for films grown on MgO and $\rm Al_2O_3$ but are typically 10$^{\circ}$ to 15$^{\circ}$ wide for films on glass. The bilayer periods determined from the low-angle peak positions are (41.0 $\pm$ 0.1)\AA\ for the MgO film and (40.9 $\pm$ 0.2)\AA\ for the glass film, the same within experimental error. $\phi$ scans of the MgO and Fe [110] peaks for the film on the MgO substrate (not shown) demonstrate that it is oriented in-plane. While $\theta$-2$\theta$ scans for (011)-oriented multilayers grown on $\rm Al_2O_3$ substrates (not shown) also show multiple high-angle superlattice satellites, the $\phi$ scans for this film indicate only weak orientation in-plane. The shape of the high-angle peaks and their superlattice satellites are described by a well-known theory.\cite{fullerton3} Application of this theory to the Fe/Si multilayers is difficult because the iron silicide lattice constant, the thickness of the remaining pure Fe and the thickness of the iron silicide spacer can be estimated only roughly. A precise determination of the silicide lattice constant should make a quantitative analysis of these satellite features possible. \section{Magnetic Characterization} Figure~\ref{bvsh}a shows magnetization curves for 60-repeat (Fe40\AA/Si14\AA) multilayers grown simultaneously on glass and MgO(001). The saturation fields H$\rm _s$ appear to be similar for the two films. On the other hand, the remanent magnetization is 58\% for the film on glass and 7\% for the film grown on MgO. A remanence as low as 1\% has been observed for other multilayers grown on MgO substrates. SQUID magnetometer data taken up to higher fields gives a saturation field of 9.75 kOe for the multilayer on MgO at room temperature. Assuming for a moment that the interlayer coupling is purely bilinear in nature, a well-known formula relates the saturation field to the AF coupling strength: A$_{12} \, = \, H_s M_s t_{Fe} / 4$ where M$\rm _s$ is the saturation magnetization and t$\rm _{Fe}$ is the thickness of an individual Fe layer.\cite{krebs} Use of this equation with H$\rm _s$ = 9.75 kOe and the measured magnetization M$_s$ = 1271 emu/cm$^3$ gives A$\rm _{12}$ = 1.2 erg/cm$^2$. This AF coupling value is comparable in size to the coupling measured in metal/metal multilayers multilayers\cite{parkin}. Figure~\ref{bvsh}b shows magnetization curves for Fe100\AA/Si14\AA/Fe100\AA trilayer films grown on Al$_2$O$_3$ (11\={2}0), Ge(001), and MgO(001) substrates. All three of the magnetization curves in Fig.~\ref{bvsh}b were taken with the field applied along an Fe(100) easy direction. Significant in-plane anisotropy of the magnetization curves occurs for the films on the Ge and MgO substrates, similar to what has been observed for Fe/Cr/Fe trilayers.\cite{krebs} The observation of magnetocrystalline anisotropy in the film on MgO but not in the film grown on Al$_2$O$_3$ is consistent with expectations from the $\phi$ scans, which show that the in-plane orientation of the film on MgO is much stronger. Figure~\ref{bvsh}b once again demonstrates that the degree of remanence in Fe/Si multilayers is strongly related to the quality of layering. While the remanent magnetization of the epitaxial trilayers on Ge and MgO is only about 5\% of the saturated value, the remanence of the polycrystalline trilayer on Al$_2$O$_3$ is close to 50\%. The remanent magnetization of the trilayers on Ge and MgO is about 5\% in the in-plane hard direction (H $\|$ Fe(110)) as well. A SQUID magnetometer has been used to measure the magnetization curves of the IBS-grown Fe/Si multilayers at lower temperatures.\cite{michel} The temperature dependence of the remanent magnetization of these films is similar to that reported by other authors.\cite{mattson} \section{Discussion} At the moment it is not possible to tell why the in-plane ordering of the films grown on Al$_2$O$_3$(11\={2}0) is inferior to that grown on the (001) MgO and Ge substrates. The difficulty with the Al$_2$O$_3$ growth may have to do with the 6$^{\circ}$ miscut of the substrates, or it may be due to an intrinsic difficulty with (011) growth of the Fe/Si multilayers. Previous work has shown that AF coupling in Fe/Si multilayers is dependent upon formation of a metastable iron silicide spacer layer phase.\cite{fesiprb} The possibility exists that the spacer silicide does not grow well on Fe in the (011) orientation. This question can be answered only by further growth studies on better (011) substrates and careful structural characterizations. A related question is whether the larger remanent magnet moment in the (011)-textured films might be due to a fundamental difference in magnetic properties from the (001)-textured films. Because a 46-repeat Fe/Si multilayer grown on Al$_2$O$_3$ has a remanence of only about 10\%, this is unlikely. Undoubtedly the trilayer on Al$_2$O$_3$ has a higher remanence than the multilayer because the thinner film is more greatly impacted by the poor substrate surface quality. The staircase morphology caused by the 6$^{\circ}$ miscut of this Al$_2$O$_3$ substrate may lead to wavy interfaces between the Fe and iron silicide films or to pinholes through the silicide layers. Wavy interfaces can cause increased magnetostatic coupling or even biquadratic coupling,\cite{slonczewski} both of which would tend to increase the remanence. The large remanence of the multilayers grown on glass substrates is likely also due to pinholes or magnetostatic coupling. Pinhole-induced may explain the unusual temperature dependence of the remanence. (Magnetostatic coupling is expected to be approximately temperature-independent.) Fe atoms in bridges through the silicide spacer layers are expected have a reduced Curie temperature. A larger remanence at low temperature therefore makes sense if the remanence is derived from pinhole coupling and is not an intrinsic effect. Low Curie-temperature material may also be present in the iron silicide spacer layer or in at the iron/iron silicide interfaces. By growing on a number of substrate materials and by using different deposition conditions, Fe/Si multilayers have been prepared with a varying degree of ordering. A large amount of accumulated evidence demonstrates that high remanence of the magnetization curves in Fe/Si multilayers is associated with interface roughness. The remanence is therefore not likely to be related to unusual exchange coupling but instead to originate from defects, perhaps pinholes through the silicide spacer layer. Since the remanent magnetization is caused by extrinsic effects, future studies should concentrate instead on measurements of the saturation field of the magnetization curves in order to learn more about the interlayer coupling. \bigskip \noindent We would like to thank E.E. Fullerton, J.A. Borchers, R.M. Osgood III and Y. Huai for helpful discussions, and B.H. O'Dell and S. Torres for technical assistance. Part of this work was performed under the auspices of the U.S. Department of Energy by LLNL under contract No. W-7405-ENG-48. \clearpage
\section{ \normalsize} \def\ds{\displaystyle\bf INTRODUCTION. } \hspace*{6mm} Among different representations of a given compact Lie group $G$ the model space ${\cal M}$ plays a distinguished role. By definition,\REF{1} the model space is a direct sum of all irreducible representations ${\cal H}_j$ with multiplicity one \begin{equation}\label} \def\eeq{\end{equation}{0.0} {\cal M} = \sum\limits_j \oplus{\cal H}_j \eeq realized in some universal way. A most popular form of ${\cal M}$ is a space of holomorphic functions on the Borel subgroup $\cal B$ of complexified form of the group $G$. In this construction the Borel subgroup is considered as an affine space. A study of model spaces provides a natural language for investigation of physical models. For example, the popular model of $2$-dimensional quantum gravity, introduced by Polyakov,\REF{2} may be interpreted in terms of the model space of Virasoro algebra.\REF{3} A finite-dimensional quantum group with deformation parameter, depending on the central charge, naturally appears in this context. In the present paper, which was written with an intent to find new applications of model space in modern mathematical physics, we discuss a $q$-analogue of the model space related to $q$-deformed Lie group $G_q$. For this purpose we introduce and examine several "coordinatizations" of the quantum space $(T^*\B)_q$. As a by-product we obtain some generating matrices for the set of Clebsch-Gordan coefficients (CGC). To our knowledge this result is new even for the non-deformed case. Throughout the paper we systematically and intentionally make use of the $R$-matrix formalism, which we believe is the most convenient and powerful tool to get explicit results in the domain of quantum groups. To avoid the known difficulties with compact forms of quantum groups we adopt here a convention to work with complexified objects (groups, algebras) and their finite dimensional representations on a formal algebraic level. We also do not discuss subtleties arising in the case of $q$ being a root of unity. Most of formulae given in this paper in $R$-matrix form have universal validity. However, the concrete results are illustrated on the simplest example $G_q=SL_q(2)$. The generalization to other groups needs more technical details such as an explicit structure of $R$-matrices and related objects. Mentioned above "coordinatizations" of $(T^*\B)_q$ arise from two possible decompositions of the matrix $L$ (in usual notations $L=L_+L_-^{-1}$, it comprises all generators of the corresponding quantum Lie algebra) : $$ L = U\,D\,U^{-1} \;\;\; {\rm and} \;\;\; L = A\,B\,A^{-1} \ , $$ where $D$ is a diagonal unimodular matrix, $U$ is a deformation of unitary matrix, $A$ and $B$ are unimodular upper and lower triangular matrices. As we shall clarify below, the matrices $A$ and $B$ admit a natural interpretation as the coordinates in the base and in the fiber of $(T^*\B)_q$, whereas entries of the matrix $U$ will be shown to provide basic shifts on the model space $\cal M$ and generate $q$-analogues of Clebsch-Gordan coefficients for the quantum group $G_q$. The explicit connection between $U$ and $(A,B)$ will be demonstrated on the example of $SL_q(2)$. It should be mentioned that an object like the matrix $U$ appeared first in Refs.4,5 (later it was used also in Ref.6), where it was interpreted as a "chiral" component of the quantum group-like element $g$. In the present paper we give another interpretation and application of the matrix $U$ in the context of a model space. Let us briefly describe the contents of the present paper. In the Sec.$\!$ II the definition of the cotangent bundle for a quantum group is reminded. Next we introduce an object of especial interest for us -- the algebra $\cal U$ generated by the entries of the matrix $U$ which diagonalizes the coordinate in a fiber of $(T^*G)_q$. We derive explicit relations for this algebra in the case of $G=SL(2)$. In the Sec.$\!$ III we consider a non-deformed limit ($q=1$) of the algebra $\cal U$ and construct an explicit representation. For the case of $G=SL(2)$ we show that the matrix $U_0$ generates Clebsch-Gordan coefficients (CGC) for the corresponding non-deformed Lie algebra. The Borel subgroup $\cal B$ and the space $T^*\B$ naturally appear here. Finally, we discuss a connection of our results with the Wigner-Eckart theorem. In the Sec.$\!$ IV we construct representations of the algebra $\cal U$ (for $q\neq 1$) for the case of $SL(2)$ in two different ways. The first one uses the language of $q$-oscillators. The second is based on explicit realization of $(T^*\B)_q$ and hence involves a notion of quantum model space. Here we show that the matrix $U$ is a "generating matrix" for CGC for deformed Lie algebra. We also give some comments on the generalized version of the Wigner-Eckart theorem. \section{\normalsize} \def\ds{\displaystyle\bf $\bf (T^*G)_q$ AND RELATED OBJECTS.} \setcounter{equation}{0} \hspace*{6mm} There exist three symplectic manifolds (from the physical point of view they are phase spaces) naturally related to a given Lie group $G$ and its Lie algebra ${\cal J}$: $\langle 1 \rangle$\ \ $T^*G$ -- the cotangent bundle for the group $G$; $\langle 2 \rangle$\ \ $T^*\B$ -- the cotangent bundle for the Borel subgroup $\B$; $\langle 3 \rangle$\ \ $\cal O$ -- an orbit of the co-adjoint action of $G$ on ${\cal J}^*$. For instance, in the case of $G=SL(2)$ (which will be our main example) these spaces are six-, four-, and two-dimensional, correspondingly. The method of geometric quantization\REF{7} provides a representation theory for $\langle 1 \rangle$, $\langle 2 \rangle$ and $\langle 3 \rangle$. Turning from classical to quantum groups, one can try to construct a representation theory for the deformed analogues of these manifolds. In the present paper we shall deal with deformations of the spaces $\langle 1 \rangle$ and $\langle 2 \rangle$. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf A. Description of $\bf (T^*G)_q$. } \hspace*{6mm} Let $G_q$ be a deformation of the Lie group $G$ and ${\cal J}_q$ be a deformation of the corresponding Lie algebra ${\cal J}$. The deformed cotangent bundle $(T^*G)_q$ is a non-commutative manifold, i.e., according to the ideology developed by A.Connes,\REF{8} its coordinates are (non-commuting) generators of some associative algebra. A point on this manifold is parameterized by the pair $(g,L)$, where $g\in G_q$ is a coordinate in the base of the bundle, and $L$ is a coordinate in a fiber. The structure of $(T^*G)_q$ is defined via commutation relations between the coordinates in the base and in a fiber. An appropriate $R$-matrix form of these relations was proposed in Ref.5: \begin{equation}\label} \def\eeq{\end{equation}{1.1} R_{\pm} \up{1}{g} \,\up{2}{g} \,=\, \up{2}{g} \, \up{1}{g} R_{\pm} \,, \eeq \begin{equation}\label} \def\eeq{\end{equation}{1.2} R_- \up{1}{g} \up{2}{L} \,=\, \up{2}{L} R_+ \up{1}{g} \,, \eeq \begin{equation}\label} \def\eeq{\end{equation}{1.3} \up{1}{L} R_-^{-1} \up{2}{L} R_- \,=\, R_+^{-1} \up{2}{L} R_+ \up{1}{L} \,. \eeq Here and below we use the formalism developed in Ref.9, i.e., objects like $g$ and $L$ are considered as matrices (say, $L\in {\cal J}} \def\Uc{{\cal U}_q\otimes V$, where $V$ stands for auxiliary space). We use the standard notations for tensor products: $\up{1}{L}\up{}{=L\otimes I}\in{\cal J}} \def\Uc{{\cal U}_q\otimes V\otimes V$, etc. Let us take the parameter $q$, which appears in the theory of quantum groups, in the following form \begin{equation}\label} \def\eeq{\end{equation}{0} q=e^{\gamma\hbar} , \eeq where $\hbar$ is the Planck constant (the parameter of quantization) and $\ga$ is the deformation parameter. In physical applications it is most natural to suppose that $\ga$ is either pure real ($q$ belongs to the real axis) or pure imaginary ($q$ belongs to the unit circle at the complex plane). The second form of $q$ is typical for the WZW theory.\REF{4,6,10} {}For $|q|=1$ we suppose also that $q$ is not a root of unity. It should be mentioned that for both variants of choice of $\ga$ in (\ref{0}) the definition of $q$-number \begin{equation}\label} \def\eeq{\end{equation}{0.2} [x] \equiv \frac{q^x - q^{-x}}{q-q^{-1}} \eeq is invariant with respect to complex conjugation of $q$, i.e. $\overline{[x]}=[\overline{x}]$. This property becomes important if one discusses involutions of deformed Lie algebras. \begin{defn}\label{DL} The algebra $\cal L$ is an associative algebra generated by entries of the matrix $L$ which obeys relation (\ref{1.3}). \end{defn} An important fact -- the connection of algebra $\cal L$ with the corresponding quantum Lie algebra ${\cal J}} \def\Uc{{\cal U}_q$ was established in Ref.11 in the following form. \begin{prop}\label{P4} Let matrices $L_+$, $L_-$ obey the following exchange relations \begin{equation}\label} \def\eeq{\end{equation}{L2} R_{\pm} \up{1}{L_+} \up{2}{L_+} = \up{2}{L_+} \up{1}{L_+} R_{\pm} ,\;\; R_{\pm} \up{1}{L_-} \up{2}{L_-} = \up{2}{L_-} \up{1}{L_-} R_{\pm}, \;\; R_{+} \up{1}{L_+} \up{2}{L_-} = \up{2}{L_-} \up{1}{L_+} R_{+} \;\; . \eeq Then the matrix $ L= L_+ \, L_{-}^{-1}$ satisfies the relation (\ref{1.3}). \end{prop} This statement implies that the algebra $\cal L$ is isomorphic (up to some technical details which we do not discuss here) to corresponding quantum Lie algebra $U_q({\cal J}} \def\Uc{{\cal U})$ [which is defined by (\ref{L2}), see, e.g., Ref.9]. Consider now the relations (\ref{1.1})-(\ref{1.3}) for $g$ and $L$ being $2\times 2$ matrices \begin{equation}\label} \def\eeq{\end{equation}{g,L} g=\left(\begin{array}} \def\er{\end{array}{cc} g_1&g_2\\g_3&g_4\er\right),\;\;\;\; L=\left(\begin{array}} \def\er{\end{array}{cc}A&B\\C&D\er\right) \eeq and the $R$-matrices taken in the form \begin{equation}\label} \def\eeq{\end{equation}{1.4} R_{+}=q^{-1/2}\left(\begin{array}} \def\er{\end{array}{cccc} q & 0 & 0 & 0 \\0 & 1 & \omega & 0 \\0 & 0 & 1 & 0 \\0 & 0 & 0 & q \er \right) ,\; \omega \equiv q-q^{-1}; \;\;\;\;\;\; R_{-}=PR_{+}^{-1}P \eeq ($P$ denotes the permutation operator: $P\!\up{1}{g}\!P\!=\,\up{2}{g}$, etc.). In this case (\ref{1.1})-(\ref{1.3}) define the cotangent bundle for the quantum group $G_q=GL_q(2)$; each of $R$-matrix equations (\ref{1.1}) and (\ref{1.3}) is equivalent to six independent relations: \begin{equation}\label} \def\eeq{\end{equation}{1.35} \begin{array}} \def\er{\end{array}{c} q\,g_1\,g_2= g_2\,g_1,\;\; q\,g_1\,g_3= g_3\,g_1,\;\; q\,g_2\,g_4= g_4\,g_2,\;\; q\,g_3\,g_4= g_4\,g_3,\;\; \\ \\ g_2\,g_3= g_3\,g_2,\;\;\;\; g_1\,g_4 - q^{-1}\, g_4\,g_1 = -\omega \,g_2\,g_3 ; \er \eeq and \begin{equation}\label} \def\eeq{\end{equation}{1.5} \begin{array}} \def\er{\end{array}{c} [A,B]=-q^{-1}\omega BD,\;\;\; [A,C]=q^{-1}\omega DC,\;\;\;[A,D]=0, \\ \\ CD=q^2DC,\;\;\;BD=q^{-2}DB,\;\;\;[B,C]=q^{-1}\omega D (D-A). \er \eeq The equation (\ref{1.2}) gives the following relations \begin{equation}\label} \def\eeq{\end{equation}{gL} \begin{array}} \def\er{\end{array}{ll} g_1 A = q A g_1 + \omega B g_3,\; & g_1 B = B g_1, \\ [1mm] g_2 A = q A g_2 + \omega B g_4,\; & g_2 B = B g_2, \\ [1mm] g_3 A = q^{-1} A g_3 + \omega g_1 C,\; & g_3 B = B g_3 +\omega g_1 D, \\ [1mm] g_4 A = q^{-1} A g_4 + \omega g_2 C,\; & g_4 B = B g_4 +\omega g_2 D, \\ [1mm] g_1 C = C g_1 + q^{-1}\omega D g_3,\; & g_1 D = q^{-1}\, D g_1, \\ [1mm] g_2 C = C g_2 + q^{-1}\omega D g_4,\; & g_2 D = q^{-1}\, D g_2, \\ [1mm] g_3 C = C g_3, & g_3 D = q D g_3, \\ [1mm] g_4 C = C g_4, & g_4 D = q D g_4. \er \eeq Next, let us recall the well-known statement (see, e.g., Ref.9): \begin{prop}\label{P1} The algebra generated by the entries of the matrix $g$ obeying (\ref{1.35}) possesses the central element (``deformed determinant'') \begin{equation}\label} \def\eeq{\end{equation}{1.65} {\det}_q g = g_1\,g_4-q^{-1}g_2\,g_3 . \eeq \end{prop} Similarly, for the algebra $\cal L$ in the case of $GL_q(2)$ one can check the following. \begin{prop}\label{P2} The algebra with generators $A$, $B$, $C$, $D$ obeying (\ref{1.5}) possesses two central elements: \begin{equation}\label} \def\eeq{\end{equation}{1.6} K_1 = qA + q^{-1} D, \;\;\; K_2 = q^{-1} AD - q BC. \eeq \end{prop} Finally, using the commutation relations (\ref{gL}), one can check that \begin{prop}\label{P3} The operators ${\det}_q g$ and $K_2$ commute with all entries of the matrices $g$ and $L$. \end{prop} This implies that, fixing values of ${\det}_q g$ and $K_2$, one gets a certain subalgebra of the algebra defined by (\ref{1.35})-(\ref{gL}). \begin{defn}\label{D1} Relations (\ref{1.35})-(\ref{gL}) for ${\det}_q g = 1$ and $ K_2 = {\rm const}$ define the cotangent bundle for the quantum group $G_q=SL_q(2)$. \end{defn} Let us underline that the above definitions and statements can be easily generalized, say to the case of $SL_q(N)$. In our case the algebra $\cal L$ is isomorphic to the quantum Lie algebra ${\cal J}_q=U_q(sl(2))$ (introduced first in Ref.12) which is defined by the relations \begin{equation}\label} \def\eeq{\end{equation}{L3} [l_+ , l_- ] = \frac{q^{2l_3}-q^{-2l_3}}{q-q^{-1}}\equiv [2l_3],\;\;\;\; q^{l_3} \, l_{\pm} = q^{\pm 1} l_{\pm} \, q^{l_3} \eeq and the matrices $L_{\pm}$ can be chosen as follows: \begin{equation}\label} \def\eeq{\end{equation}{L4} L_+ = \left( \begin{array}} \def\er{\end{array}{cc} q^{l_3} & \omega q^{1/2} l_- \\ 0 & q^{-l_3} \er \right),\;\;\;\; L_- = \left( \begin{array}} \def\er{\end{array}{cc} q^{-l_3} & 0 \\ -\omega q^{-1/2} l_+ & q^{l_3} \er \right) . \eeq Note that the matrix $L$ in the Proposition \ref{P4} is defined only up to a scaling factor. Thus, for $L_+$, $L_-$ given in (\ref{L4}), we may choose $L$ as follows \begin{equation}\label} \def\eeq{\end{equation}{L5} L= q^2 L_+ L_{-}^{-1} = \left(\begin{array}} \def\er{\end{array}{cc} q C - q^{-2 l_3} & q^{5/2} \omega l_- q^{-l_3} \\ q^{-1/2}\omega l_+ q^{-l_3} & q^2 q^{-2l_3} \er \right). \eeq Here $C$ stands for the Casimir operator of $U_q(sl(2))$ : \begin{equation}\label} \def\eeq{\end{equation}{L8} C = {\omega}^2 l_- l_+ + q^{2l_3+1} + q^{-(2l_3+1)} = q^{2\wh{j}+1} + q^{-(2\wh{j}+1)} , \eeq where $\wh{j}$ is the operator of spin. According to Proposition \ref{P4}, the matrix (\ref{L5}) satisfies (\ref{1.3}). Therefore, it provides a (fundamental) representation of the algebra ${\cal L}$ for $U_q(sl(2))$. In this representation the central elements (\ref{1.6}) are given by \begin{equation}\label} \def\eeq{\end{equation}{L6} K_1 = q^2 C, \;\;\; K_2 = q^3, \eeq Note that the scaling factor $q^{2}$ introduced in (\ref{L5}) has changed the values of $K_1$ and $K_2$. The choice of such normalization in (\ref{L5}) will be explained later. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf B. Connection with quantum $\bf 6j$-symbols. } \hspace*{6mm} Let us remind the theorem which describes an important property of the algebra ${\cal L}$ for $U_q(sl(2))$ (this statement appeared first in Ref.5). \begin{theorem} Let $D\equiv D(p)$ be the unimodular diagonal matrix \begin{equation}\label} \def\eeq{\end{equation}{1.75} D = \left( \begin{array}} \def\er{\end{array}{cc} q^{p/\hbar} & \\ & q^{-p/\hbar} \er \right) , \eeq and let 2$\times$2 matrix $U$ satisfy the following exchange relations \begin{equation}\label} \def\eeq{\end{equation}{1.8} \up{1}{D}\, \up{2}{U} = \up{2}{U} \up{1}{D} \sigma,\;\; \; \up{2}{D}\, \up{1}{U} = \up{1}{U} \up{2}{D} \sigma, \;\;\; \sigma= {\rm diag} (q^{-1},q,q,q^{-1}), \eeq \begin{equation}\label} \def\eeq{\end{equation}{1.9} R_+ \up{1}{U} \, \up{2}{U} = \up{2}{U} \,\up{1}{U} {\cal R}_+ (p), \;\; R_- \up{1}{U}\, \up{2}{U} = \up{2}{U}\, \up{1}{U} {\cal R}_- (p), \eeq where $R_{\pm}$ are the standard $R$-matrices (\ref{1.4}) and \begin{equation}\label} \def\eeq{\end{equation}{1.10} {\cal R}_+ (p)= P{\cal R}_-^{-1} (p) P = q^{-1/2} \left( \begin{array}} \def\er{\end{array}{cccc} q & & & \\ & \ds \frac{\sqrt{ [p/\hbar +1] [p/\hbar -1] }}{[p/\hbar]} & \ds \frac{q^{p/\hbar}}{[p/\hbar]} & \\ & & \\ & \ds -\frac{q^{-p/\hbar}}{[p/\hbar]} & \ds \frac{\sqrt{ [p/\hbar +1] [p/\hbar -1] }}{[p/\hbar]} & \\ & & & q \er \right), \eeq (here $[x]$ denotes a "q-number" (\ref{0.2})). Then matrix $L$ constructed by means of the similarity transformation \begin{equation}\label} \def\eeq{\end{equation}{1.7} L = U D\, U^{-1}, \;\;\; \eeq satisfies the relation (\ref{1.3}) and therefore its entries generate an algebra ${\cal L}$ isomorphic to $U_q(sl(2))$. \end{theorem} The proof is given in appendix A. It makes use of the identity \begin{equation}\label} \def\eeq{\end{equation}{RD} {\cal R}_- (p) = (\up{1}{D})^{-1} {\cal R}_+ (p) \sigma \up{1}{D} . \eeq {\it Remark:} \ } \def\proof{{\it Proof}\ A consequence of (\ref{1.8}) is the commutativity of $L$ and $D$ \begin{equation}\label} \def\eeq{\end{equation}{1.12} \up{1}{L} \, \up{2}{D} = \up{2}{D} \, \up{1}{L}, \eeq which implies that $p$ commutes with all elements of $\cal L$. Later we shall interpret $p$ as the operator of spin. {\it Remark:} \ } \def\proof{{\it Proof}\ Properly generalizing objects which enter the Theorem 1, one can extend this theorem to the case of any quantum semisimple Lie algebra.\REF{13} In particular, the matrix $D$ for $U_q(sl(N))$ is found to be : $D(\vec{p}\,)\,=\, const\cdot q^{\vec{H}\,\o\,\ds\vec{p} }$,\ where $\vec{p}$ consists of the operators corresponding to components of the weight vector (i.e., on each irreducible representation they are multiples of unity) and $\vec{H}$ consists of the generators $H_i$ of the Cartan subalgebra. An explicit form of ${\cal R}(p)$ for $U_q(sl(N))$ was obtained in Ref.14. {\it Remark:} \ } \def\proof{{\it Proof}\ The matrix ${\cal R}(p)$ obeys the deformed Yang-Baxter equation,\REF{14-16,5} which can be written, for example, as follows : \begin{equation}\label} \def\eeq{\end{equation}{1.13} \up{1}{Q} \up{23}{\cal R}_+ (p) (\up{1}{Q})^{-1} \up{13}{\cal R}_+ (p) \up{3}{Q} \up{12}{\cal R}_+ (p) (\up{3}{Q})^{-1} = \up{12}{\cal R}_+ (p) \up{2}{Q} \up{13}{\cal R}_+ (p) (\up{2}{Q})^{-1} \up{23}{\cal R}_+ (p), \eeq where for ${\cal J}} \def\Uc{{\cal U}_q=U_q(sl(2))$ the matrix $Q = \biggl(\! \begin{array}} \def\er{\end{array}{cc} e^{i\xi} & \\ &\! e^{-i\xi} \er\!\biggr)$ contains an extra variable $\xi$, conjugated with $p$: \begin{equation}\label} \def\eeq{\end{equation}{1.15} [p,\xi] = -i \hbar, \;\;\;\;\; q^{p/\hbar}\,e^{i\xi} = q\,e^{i\xi}\,q^{p/\hbar}. \eeq This variable $\xi$ belongs to the algebra $\cal U$ but does not enter matrix $L$. An explicit expression for $\xi$ will be given below. The general form of $Q$ for $U_q(sl(N))$ can be easily found\REF{13} : $ Q\,=\,e^{ i\,\vec{H}\,\o\,\ds\vec{\xi} }$, where components of $\xi_i$ are operators conjugated to $p_i$ : $[p_j,\xi_k]=-i\hbar\,\dl_{jk}$. The matrix ${\cal R}(p)$ was discussed in physical literature in different contexts. In particular, it plays significant role in studies of quantum Liouville\REF{15,16} and WZW\REF{4-6} models; its relation to Calogero-Moser model was recently discussed in Ref.17. But for us more important fact is a connection of ${\cal R}(p)$ with the quantum $6j$-symbols: the entries of (\ref{1.10}) calculated on irreducible representations coincide (up to some normalization) with the values of some $6j$-symbols for $U_q(sl(2))$ (exact formulae are given in Ref.18, generalizations are discussed in Ref.13). This connection allows to assume that objects like the matrix $U$ should be interpreted in terms of Clebsch-Gordan coefficients (CGC). Below we demonstrate that $U$ is indeed a ``generating matrix'' for CGC and clarify its relation to $(T^*\B)_q$. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf C. Algebra ${\cal U}$. } \begin{defn}\label{DU} The algebra $\cal U$ is an associative algebra generated by entries of matrix $U = \left( \begin{array}} \def\er{\end{array}{cc} U_1 & U_2 \\ U_3 & U_4 \er\right)$ and the operator $p$ such that relations (\ref{1.75})-(\ref{1.10}) hold. \end{defn} {\it Remark:} \ } \def\proof{{\it Proof}\ For simplicity we restricted our consideration to the case of $\cal U$ associated with $U_q(sl(2))$. Let us stress that the case of $\cal U$ associated with $U_q(sl(N))$ can be studied similarly but it will involve more technical details. On the other hand, it might be rather cumbrous to obtain exact formulae for $\cal U$ associated with $U_q({\cal J}} \def\Uc{{\cal U})$ in the case of ${\cal J}} \def\Uc{{\cal U}$ being generic semisimple Lie algebra. Let us give an explicit form of the defining relations (\ref{1.9}) : \begin{equation}\label} \def\eeq{\end{equation}{2.1} U_1 U_3 = q^{-1} U_3 U_1, \;\;\;\;\; U_2 U_4 = q^{-1} U_4 U_2, \eeq \begin{equation}\label} \def\eeq{\end{equation}{2.2} U_1 U_2 = U_2 U_1 \sqrt{\frac{[p/\hbar -1]}{[p/\hbar +1]}}, \;\;\; U_3 U_4 = U_4 U_3 \sqrt{\frac{[p/\hbar -1]}{[p/\hbar +1]}} \eeq \begin{equation}\label} \def\eeq{\end{equation}{2.3} U_1 U_4 = U_4 U_1 \frac{\sqrt{[p/\hbar +1] [p/\hbar -1]}} {[p/\hbar]} - U_3 U_2 \frac{q^{p/\hbar}}{[p/\hbar ]} \eeq \begin{equation}\label} \def\eeq{\end{equation}{2.4} U_3 U_2 = U_2 U_3 \frac{\sqrt{[p/\hbar +1] [p/\hbar -1]}}{[p/\hbar]} - U_1 U_4 \frac{q^{-p/\hbar}}{[p/\hbar ]} \eeq The rest of the relations contained in (\ref{1.9}) are not independent and can be deduced from (\ref{2.1})-(\ref{2.4}). Additionally, from (\ref{1.8}) one gets \begin{equation}\label} \def\eeq{\end{equation}{2.5} \begin{array}} \def\er{\end{array}{ll} q^{p/\hbar} \, U_1 = q^{-1} \, U_1 \, q^{p/\hbar}, \; & q^{p/\hbar} \, U_2 = q \, U_2 \, q^{p/\hbar},\\ & \\ q^{p/\hbar} \, U_3 = q^{-1} \, U_3 \, q^{p/\hbar},\;& q^{p/\hbar} \, U_4 = q \, U_4 \, q^{p/\hbar}.\er \eeq Thus, relations (\ref{2.1})-(\ref{2.5}) describe the algebra ${\cal U}$. Using them, one may verify the following statement: \begin{prop} A central element of ${\cal U}$ is given by the "deformed" determinant of the matrix $U$: \begin{equation}\label} \def\eeq{\end{equation}{2.6} {\rm Det} U \equiv U_1 U_4 \sqrt{\frac{[p/\hbar +1]}{[p/\hbar ]}} - U_2 U_3 \sqrt{\frac{[p/\hbar -1]}{[p/\hbar]}}= q U_4 U_1 \sqrt{\frac{[p/\hbar -1]}{[p/\hbar]}} - q U_3 U_2 \sqrt{\frac{[p/\hbar +1]}{[p/\hbar ]}} . \eeq \end{prop} For fixed value of ${\rm Det} U$ the algebra ${\cal U}$ contains only four independent generators. In classical limit $(\hbar = 0)$ they become the coordinates on 4-dimensional phase space. For further discussion it is convenient to introduce new variables instead of $U_i$: \begin{equation}\label} \def\eeq{\end{equation}{2.7} \wh{U}_i = U_i \sqrt{[p/\hbar]}. \eeq The coordinates $\{p$, $\wh{U}_i\}$ form a new set of generators of the algebra ${\cal U}$. The commutation relations (\ref{2.1})-(\ref{2.5}) rewritten in terms of the new generators acquire a simpler form: \begin{equation}\label} \def\eeq{\end{equation}{2.8} \wh{U}_1 \wh{U}_3 = q^{-1} \wh{U}_3 \wh{U}_1, \;\; \wh{U}_2 \wh{U}_4 = q^{-1} \wh{U}_4 \wh{U}_2, \;\; \wh{U}_1 \wh{U}_2 = \wh{U}_2 \wh{U}_1, \;\; \wh{U}_3 \wh{U}_4 = \wh{U}_4 \wh{U}_3 \eeq \begin{equation}\label} \def\eeq{\end{equation}{2.10} \wh{U}_1 \wh{U}_4 = \wh{U}_4 \wh{U}_1 \frac{[p/\hbar +1]}{[p/\hbar]} - \wh{U}_3 \wh{U}_2 \frac{q^{p/\hbar}}{[p/\hbar ]}, \eeq \begin{equation}\label} \def\eeq{\end{equation}{2.11} \wh{U}_3 \wh{U}_2 = \wh{U}_2 \wh{U}_3 \frac{[p/\hbar +1] }{[p/\hbar]} - \wh{U}_1 \wh{U}_4 \frac{q^{-p/\hbar}}{[p/\hbar ]}, \eeq \vspace*{1mm} \begin{equation}\label} \def\eeq{\end{equation}{2.9} \begin{array}} \def\er{\end{array}{ll} q^{p/\hbar} \, {\wh U}_1 = q^{-1} \, {\wh U}_1 \, q^{p/\hbar}, \;& q^{p/\hbar} \, {\wh U}_2 = q \, {\wh U}_2 \, q^{p/\hbar}, \\ & \\ q^{p/\hbar} \, {\wh U}_3 = q^{-1} \, {\wh U}_3 \, q^{p/\hbar},\;& q^{p/\hbar} \, {\wh U}_4 = q \, {\wh U}_4 \, q^{p/\hbar}. \er \eeq The central element (\ref{2.6}) in new variables looks as follows \begin{equation}\label} \def\eeq{\end{equation}{2.12} {\rm Det} U \equiv ( \wh{U}_1 \wh{U}_4 - \wh{U}_2 \wh{U}_3)\frac{1} {[p/\hbar ]} = ( \wh{U}_4 \wh{U}_1 - \wh{U}_3 \wh{U}_2 )\frac{q}{[p/\hbar]}. \eeq The explicit form of the matrix inverse to $\wh{U}$, which we shall need later, is \begin{equation}\label} \def\eeq{\end{equation}{2.125} \wh{U}^{-1} = \frac{1}{{\rm Det}U} \left(\begin{array}} \def\er{\end{array}{cc} \wh{U}_4 & -q \wh{U}_2 \\ - \wh{U}_3 & q \wh{U}_1 \er\right)\frac{1}{[p/\hbar]}. \eeq Finally, from (\ref{2.7}) we conclude that the expression (\ref{1.7}) for the matrix $L$ looks similarly in terms of new matrix $\wh{U}$: \begin{equation}\label} \def\eeq{\end{equation}{2.130} L \,=\, U\,D\,U^{-1} \,=\, \wh{U} \, D \, \wh{U}^{-1} \ . \eeq \section{ \normalsize} \def\ds{\displaystyle\bf NON-DEFORMED CASE. } \setcounter{equation}{0} \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf A. Representation of algebra ${\cal U}_0$. } \hspace*{6mm} First, we consider the limit $\gamma\rightarrow 0$, $\hbar \neq 0$ (note that $q$-numbers turn into ordinary numbers),i.e., here we deal with a well understood situation -- the representation theory of $SL(2)$. An investigation of this simple non-deformed case will make further results more transparent. Let us denote the corresponding limit algebra as ${\cal U}_0$. The defining $R$-matrix relations (\ref{1.9}) now degenerate to \begin{equation}\label} \def\eeq{\end{equation}{2.12b} \up{1}{U_0}\, \up{2}{U_0} = \up{2}{U_0} \, \up{1}{U_0} {\cal R}_{\pm}^0 (p), \eeq where \begin{equation}\label} \def\eeq{\end{equation}{1.12c} {\cal R}_{+}^0 (p) = {\cal R}_{-}^0 (p) = \left( \begin{array}} \def\er{\end{array}{cccc} 1 & & & \\ & \ds\frac{\sqrt{ (p/\hbar +1) (p/\hbar -1) }}{(p/\hbar)} & \ds\frac{\hbar}{p} & \\ & & \\ & -\ds\frac{\hbar}{p} & \ds\frac{\sqrt{(p/\hbar +1) (p/\hbar -1)}}{(p/\hbar)} & \\ & & & 1 \er\right). \eeq The analogues of relations (\ref{2.8})-(\ref{2.9}) for ${\cal U}_0$ are (from now on we omit the index $0$ for the generators of ${\cal U}_0$) \begin{equation}\label} \def\eeq{\end{equation}{2.13a} p \wh{U}_1 = \wh{U}_1 (p-\hbar), \;\; p \wh{U}_2 = \wh{U}_2 (p+\hbar), \;\; p \wh{U}_3 = \wh{U}_3 (p-\hbar), \;\; p \wh{U}_4 = \wh{U}_4 (p+\hbar) , \eeq \begin{equation}\label} \def\eeq{\end{equation}{2.13b} [\wh{U}_1, \wh{U}_2 ] \, = \, [\wh{U}_1, \wh{U}_3 ] \, = \, [\wh{U}_2, \wh{U}_4 ] \, = \, [\wh{U}_3, \wh{U}_4 ] = 0 , \eeq \begin{equation}\label} \def\eeq{\end{equation}{2.13c} [ \wh{U}_1, \wh{U}_4 ] = {\rm Det}U_0 ,\;\;\;\;\; [\wh{U}_3, \wh{U}_2 ] = - {\rm Det}U_0 , \eeq where ${\rm Det} U_0$ stands for a limit version of (\ref{2.12}): \begin{equation}\label} \def\eeq{\end{equation}{2.13d} {\rm Det} U_0 = ( \wh{U}_1 \wh{U}_4 - \wh{U}_2 \wh{U}_3 ) \frac{\hbar}{p} = ( \wh{U}_4 \wh{U}_1 - \wh{U}_3 \wh{U}_2 ) \frac{\hbar}{p} . \eeq \begin{prop}\label{P5} A possible solution for (\ref{2.13a})-(\ref{2.13d}) is \begin{equation}\label} \def\eeq{\end{equation}{2.14} \wh{U}_1 = \partial_1, \;\; \wh{U}_2 = z_2, \;\; \wh{U}_3 = -\partial_2, \;\; \wh{U}_4 = z_1; \eeq \begin{equation}\label} \def\eeq{\end{equation}{2.15} p = \hbar ( z_1 \partial_1 + z_2 \partial_2 +1), \eeq where we denote $\partial_i \equiv \frac{\partial}{\partial z_i}$. \end{prop} {\it Remark:} \ } \def\proof{{\it Proof}\ The representation given by (\ref{2.14})-(\ref{2.15}) is not unique. In particular, the rescaling $\wh{U}_i\rightarrow} \def\lar{\leftarrow} \def\ov{\overline c_i \wh{U}_i$ (where $c_i$ are numerical constants such that $c_1 c_4 = c_2 c_3$) is allowable. The Proposition \ref{P5} together with the connection formula (\ref{2.7}) allows us to write out the explicit form of the matrix $U_0$ \begin{equation}\label} \def\eeq{\end{equation}{2.16} U_0 = \left( \begin{array}} \def\er{\end{array}{cc} \partial_1 & z_2 \\ -\partial_2 & z_1 \er \right)\, \sqrt{\frac{\hbar}{p}} . \eeq Note that this matrix is ``unimodular'', i.e., ${\rm Det} U_0 = ( \partial_1 z_1 + z_2 \partial_2 )\frac{\hbar}{p} = 1$. To describe the obtained representation of the algebra ${\cal U}_0$ completely one has to define a space where operators (\ref{2.14})-(\ref{2.16}) act. It is natural to think that this space is $D(z_1,z_2)$ -- a space of holomorphic functions of two complex variables. Let us recall that $D(z_1,z_2)$ is a space spanned on the vectors \begin{equation}\label} \def\eeq{\end{equation}{3.2} |j,m\rangle =\frac{z_1^{j+m}z_2^{j-m}}{\sqrt{(j+m)!(j-m)!}},\;\;\;\; j=\begin{array}} \def\er{\end{array}{lllll}0,&\!\!\frac{1}{2},&\!\!1,&\!\!\frac{3}{2},&\!\! ...\er \;\;\;\;\;m=-j,..,j \eeq and equipped with the scalar product \begin{equation}\label} \def\eeq{\end{equation}{3.3} \langle f,g\rangle =\frac{1}{(2\pi i)^2}\int \overline{f(z_1, z_2)} g(z_1,z_2) e^{-z_1\bar z_1-z_2\bar z_2}dz_1 d\bar z_1 dz_2 d\bar z_2. \eeq The system (\ref{3.2}) is orthonormal with respect to the scalar product (\ref{3.3}), that is $ <j,m|j^{\prime},m^{\prime}>=\delta_{j j^{\prime}} \delta_{m m^{\prime}}$. For the given scalar product a rule of conjugation of operators looks as follows \begin{equation}\label} \def\eeq{\end{equation}{3.5} (z_i)^{*}=\partial _i,\;\;\;(\partial _i)^{*}=z_i. \eeq The question concerning unitarity of the matrix $U_0$ is discussed in Appendix B. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf B. Connection with $\bf T^*{\cal B}$. } \hspace*{6mm} The generators of $sl(2)$ can be realized on $D(z_1,z_2)$ as differential operators: \begin{equation}\label} \def\eeq{\end{equation}{3.1} l_{+}=z_1\partial _2,\;\;l_{-}=z_2\partial _1,\;\;l_3=\begin{array}} \def\er{\end{array}{c} \frac 12 \er \! (z_1\partial _1-z_2\partial _2). \eeq Using these expressions we can compare the representation of the algebra ${\cal L}$ (or, more precisely, its limit version ${\cal L}_0$) given by Theorem 1 with the representation given by Proposition \ref{P4}. Indeed, in the limit $\ga\rightarrow} \def\lar{\leftarrow} \def\ov{\overline 0$ the initial formula (\ref{1.7}) acquires form \begin{equation}\label} \def\eeq{\end{equation}{3.05} L=I+\ga L_0 +O(\ga^2) , \;\;\; L_0 = U_0\left(\begin{array}} \def\er{\end{array}{cc} p&\\ &-p\er\right) U_0^{-1} . \eeq Substituting here the explicit expressions (\ref{2.15})-(\ref{2.16}) for $p$, $U_0$ and using the representation (\ref{3.1}) for generators of $sl(2)$, one derives the following limit form of the $L$-operator: \begin{equation}\label} \def\eeq{\end{equation}{3.15} L_0=\hbar \ \left(\begin{array}} \def\er{\end{array}{cc} 2 + z_1\,\partial_1 - z_2\,\partial_2 & 2 z_2\,\partial_1 \\ 2 z_1\,\partial_2 & 2 - z_1\,\partial_1 + z_2\, \partial_2 \er \right) = 2\hbar \ \left(\begin{array}} \def\er{\end{array}{cc} 1+l_3 & l_{-} \\ l_{+} & 1-l_3 \er \right) . \eeq Notice that (\ref{3.15}) exactly coincides with (\ref{L5}) taken in the limit $\ga \rightarrow} \def\lar{\leftarrow} \def\ov{\overline 0$. This explains why we had to introduce the factor $q^2$ in (\ref{L5}). The next observation concerning the limit of $L$-operator reads as follows. \begin{prop}\label{P6} The matrix $L_0$ in the representation (\ref{3.15}) admits the decomposition \begin{equation}\label} \def\eeq{\end{equation}{3.35a} L_0 = A_0\,B_0\,A_0^{-1}, \eeq where \begin{equation}\label} \def\eeq{\end{equation}{3.35b} A_0 = \left(\begin{array}} \def\er{\end{array}{cc} z_1^{-1/2} & -z_1^{-1/2} z_2 \\ 0 & z_1^{1/2} \er\right), \;\;\; B_0= \hbar\left(\begin{array}} \def\er{\end{array}{cc} p/\hbar+1/2 & 0 \\ 2\,\partial_2 & -(p/\hbar-1/2)\er\right) \eeq and $p$ is defined as in (\ref{2.15}). \end{prop} This statement can be verified directly. Let us comment on the meaning of this proposition. First, note that $A_0$ is a realization of a group-like element of the Borel subgroup of $SL(2)$. Moreover, this explicit form of $A_0$ is straightly connected with the construction of the model space $\cal M$ developed by Gelfand et al.\REF{1} Indeed, the space $D(z_1,z_2)$ being a realization of the model space for $SL(2)$ (compare (\ref{0.0}) and (\ref{3.2})) is spanned on monomials with arguments which are combinations of the entries of $A_0$. On the other hand, $B_0$ is of opposite (with respect to $A_0$) triangularity and its entries are operators acting on a given realization of the model space. Therefore, $B_0$ can be regarded as an element of the space dual to the corresponding Borel subalgebra. Thus, $A_0$ and $B_0$ are coordinates in the base and in a fiber of the cotangent bundle $T^*\B$. At this stage the appearance of $T^*\B$ "inside" the algebra $\cal L$ looks somewhat mysterious, but we shall clarify it later. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf C. Clebsch-Gordan coefficients. } \hspace*{6mm} Let us consider an action of the generators of the algebra ${\cal U}_0$ defined in (\ref{2.15})-(\ref{2.16}) on the space $D(z_1,z_2)$ (which is a realization of the model space). The action of these operators on the basic vectors (\ref{3.2}) is given by \begin{equation}\label} \def\eeq{\end{equation}{3.8} p \; |j,m\rangle =(2j+1)\hbar \; |j,m\rangle , \eeq \begin{equation}\label} \def\eeq{\end{equation}{3.9} \!\!\!\! \begin{array}} \def\er{\end{array}{ll} U_1|j,m\rangle =\left(\frac{j+m}{2j+1}\right)^{1/2}|j-\frac 12 ,m-\frac 12 \rangle , & U_2|j,m\rangle =\left(\frac{j-m+1}{2j+1}\right)^{1/2}|j+\frac 12 ,m-\frac 12 \rangle ,\\ \\ U_3|j,m\rangle =-\left(\frac{j-m}{2j+1}\right)^{1/2}|j-\frac 12 ,m+\frac 12 \rangle , & U_4|j,m\rangle =\left(\frac{j+m+1}{2j+1}\right)^{1/2}|j+\frac 12 ,m+\frac 12 \rangle . \er \eeq Formula (\ref{3.8}) allows us to identify the operator $p$ as $p=2\wh{j}+1$, where $\wh{j}$ is the operator of spin. Hence, invariant subspaces of $p$ on the model space are those with fixed value of spin $j$. Formulae (\ref{3.9}) show that $U_i$ are generators of the basic shifts on the model space (as illustrated on Fig.1). This observation is very important. As we shall see later, the same picture holds for $q\neq 1$. \vspace*{1mm} \hbox{ \begin{picture}(200,150)(-100,0) \put(20,20){\vector(1,0){150}} \put(70,20){\vector(0,1){131}} \multiput(90,20)(0,10){12}{\line(0,1){5}} \multiput(110,20)(0,10){12}{\line(0,1){5}} \multiput(130,20)(0,10){12}{\line(0,1){5}} \multiput(70,80)(10,0){8}{\line(1,0){5}} \multiput(70,100)(10,0){8}{\line(1,0){5}} \multiput(70,120)(10,0){8}{\line(1,0){5}} \put(73,147){$j$} \put(173,23){$m$} \put(72,5){$m\!-\!\frac{1}{2}$} \put(106,5){$m$} \put(122,5){$m\!+\!\frac{1}{2}$} \put(42,78){$j\!-\!\frac{1}{2}$} \put(42,98){$j$} \put(42,118){$j\!+\!\frac{1}{2}$} \put(75,65){$U_1$} \put(75,125){$U_2$} \put(135,65){$U_3$} \put(135,125){$U_4$} \thicklines \put(110,100){\vector(1,1){20}} \put(110,100){\vector(-1,1){20}} \put(110,100){\vector(1,-1){20}} \put(110,100){\vector(-1,-1){20}} \end{picture} } \vspace*{0.5mm} \begin{center} \parbox{9cm}{\small {\bf Fig.1:} \ Action of the operators $U_i$ on the model space.} \end{center} Now comparing the matrix elements $\langle\,j'',m''|U_i|j,m\,\rangle$ following from (\ref{3.9}) with values of the Clebsch-Gordan coefficients (CGC) for decomposition of the tensor product of irreducible representations $V_j\otimes V_{\frac 12}$ for $sl(2)$ which are given by the Van-der-Waerden formula \begin{equation}\label} \def\eeq{\end{equation}{3.10} \left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime } \\ m & m^{\prime } & m^{\prime \prime } \er \right\} = \delta_{m^{\prime \prime }, m+m^{\prime }} \sqrt{ \frac{(j+\frac 12-j^{\prime \prime })! (j+j^{\prime \prime }-\frac 12 )!(j^{\prime \prime }+\frac 12-j)! } {(j+j^{\prime \prime }+\frac 32)!} } \; \times \eeq \[ \!\! \times \sum\limits_{r\geq 0} \frac{(-1)^r \sqrt{ (j+m)!(j-m)!(j^{\prime \prime}+m^{\prime \prime})! (j^{\prime\prime }-m^{\prime \prime })!(2j^{\prime\prime }+1) } } {r!(j+\frac 12-j^{\prime \prime}-r)!(j-m-r)!(\frac 12+m^{\prime }-r)! (j^{\prime \prime }-\frac 12+m+r)!(j^{\prime \prime }-j-m^{\prime }+r)!}, \] we establish the following correspondence \begin{equation}\label} \def\eeq{\end{equation}{3.14} \! \begin{array}} \def\er{\end{array}{ll} \langle j^{\prime \prime} , m^{\prime \prime} \mid U_1\mid j, m\rangle= \delta_{j^{\prime \prime },j- \frac 12 } \left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime} \\ m & -\frac 12 & m^{\prime \prime} \er \right\} , \\ [1mm] \langle j^{\prime \prime} , m^{\prime \prime} \mid U_2 \mid j, m\rangle= \delta_{j^{\prime \prime },j+ \frac 12 }\left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime} \\ m & -\frac 12 & m^{\prime \prime} \er \right\} , \\ [1mm] \langle j^{\prime \prime} , m^{\prime \prime} \mid U_3 \mid j, m\rangle= \delta_{j^{\prime \prime },j- \frac 12 }\left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime} \\ m & \frac 12 & m^{\prime \prime} \er \right\}, \\ [1mm] \langle j^{\prime \prime} , m^{\prime \prime} \mid U_4 \mid j, m\rangle= \delta_{j^{\prime \prime },j+ \frac 12 }\left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime} \\ m & \frac 12 & m^{\prime \prime}\er \right\}. \er \eeq Thus, we proved the following statement: \begin{prop}\label{P7} The generators $U_i$ of the algebra ${\cal U}_0$ are operators of the basic shifts on the model space for $sl(2)$ and they generate the Clebsch-Gordan coefficients corresponding to decomposition of the product $V_j\otimes V_{\frac 12}$ of the irreps of sl(2). \end{prop} This statement allows to call the matrix $U_0$ a ''generating matrix'' (by analogy with the notion of a generating function) for CGC. {\it Remark:} \ } \def\proof{{\it Proof}\ Usually, introducing a generating object (well-known examples are the generating functions for different sets of polynomials, e.g., for the Legendre polynomials), one makes properties of the objects under consideration more evident. We think that the notion of generating matrix will be useful for calculations involving CGC of classical and quantum algebras. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf D. Wigner-Eckart theorem.} \hspace*{6mm} One should underline a connection of the results obtained above (Proposition \ref{P7}) and the well-known mathematical construction -- Wigner-Eckart theorem,\REF{19} which has important applications in quantum mechanics. Let us remind that the Wigner-Eckart theorem gives CGC for classical Lie algebra ${\cal J}$ as matrix elements of some set of operators. These operators are called {\it tensor operators}. They map the corresponding model space $\cal M$ onto itself and have special transformation properties under adjoint action of the algebra. In the case of ${\cal J}=sl(2)$ the Wigner-Eckart theorem reads as follows. \begin{theorem} \label{Twe} Let $l_{+}$, $l_{-}$ and $l_{3}$ be the generators of $sl(2)$ and let $T^{j\;}_{\;m}$, $m=-j,..,j$ be a set of operators acting on $\cal M$ and obeying the commutation relations \begin{equation}\label} \def\eeq{\end{equation}{we1} [l_3,T^{j\;}_{\;m}]=m\,T^{j\;}_{\;m},\;\;\;\; [l_{\pm},T^{j\;}_{\;m}]=\sqrt{(j\mp m)(j\pm m+1)}\; T^{j\;}_{\;m\pm 1}, \eeq where $j(j+1)$ is an eigenvalue of the Casimir operator for $sl(2)$. Then the matrix elements of $T^{j\;}_{\;m}$ on $\cal M$ are proportional to Clebsch-Gordan coefficients: \[ <j^{\prime\prime} m^{\prime\prime} |T^{j\;}_{\;m}|j^{\prime} m^{\prime} > = C_{j^{\prime\prime}j^{\prime}}^{j}\;\left\{\begin{array}} \def\er{\end{array}{ccc}j^{\prime} & j & j^{\prime\prime}\\ m^{\prime} & m & m^{\prime\prime} \er \right\}, \] where the coefficients $C_{j^{\prime\prime}j^{\prime}}^j$ do not depend on $m$, $m^{\prime}$, $m^{\prime\prime}$. \end{theorem} Proposition \ref{P7} says that any tensor operators of spin $j=1/2$ (that is $\{T^{1/2}_{1/2},\, T^{1/2}_{-1/2}\}$,\ $T^{1/2}_m\, :\, V_j \mapsto V_j\otimes V_{1/2}=V_{j+1/2}\oplus V_{j-1/2}$) may be constructed via the operators $U_i$ (in fact, it is evident from Fig.1). Indeed, comparing the commutation relations obtained directly from (\ref{2.16}) and (\ref{3.1}) \begin{equation}\label} \def\eeq{\end{equation}{we20} \begin{array}} \def\er{\end{array}{llll} {}[l_{+},U_1]=U_3, & [l_{+},U_2]=U_4, & [l_{+},U_3]=0, & [l_{+},U_4]=0,\\ & & & \\ {} [l_{-},U_1]=0, & [l_{-},U_2]=0, & [l_{-},U_3]=U_1, & [l_{-},U_4] = U_2,\\ & & & \\ {}[l_{3},U_1]=-\frac{1}{2} U_1, &[l_{3},U_2] =-\frac{1}{2} U_2, & [l_{3},U_3] = \frac{1}{2} U_3, & [l_{3},U_4]= \frac{1}{2} U_4 \er \eeq with Theorem \ref{Twe}, we get the following. \begin{prop}\label{P8} The generators $U_i$ of the algebra ${\cal U}_0$ form a basis for tensor operators of spin $1/2$, that is components $T^{1/2}_{1/2}$ and $T^{1/2}_{-1/2}$ of any tensor operator of spin 1/2 can be realized as linear combinations of $U_i$: \begin{equation}\label} \def\eeq{\end{equation}{3.55} T^{1/2}_{-1/2}= \mu(p) \;U_1+\nu(p) \;U_2,\;\;\; T^{1/2}_{1/2}=\mu(p) \;U_3+\nu(p) \;U_4, \eeq where $\mu(p)$ and $\nu(p)$ are functions only of $p=2\wh{j}+1$. \end{prop} \section{ \normalsize} \def\ds{\displaystyle\bf DEFORMED CASE. } \setcounter{equation}{0} \hspace*{6mm} Now we want to extend the results obtained in the previous section to the case of $q\neq 1$. In particular, we are going to examine the representations of the algebra ${\cal U}$ (see Definition \ref{DU} above) and to show that the corresponding matrix $U$ generates Clebsch-Gordan coefficients for the deformed Lie algebra. For these purposes we shall exploit a natural connection of ${\cal U}$ with $(T^*\B)_q$. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf A. The $\bf q$-oscillators approach. } \hspace*{6mm} There exist different ways to obtain desirable representations of the algebra ${\cal U}$. First we describe a more direct but less instructive method, which is similar to that used in the non-deformed case. By analogy with the non-deformed case studied above, one can assume that the entries of the matrix $U$ might be realized as operators (deformations of those obtained in Proposition \ref{P5}) acting on the space of two complex variables. Indeed, using the definition (\ref{2.6}) of the central element of $\cal U$ and taking into account the identity for $q$-numbers \begin{equation}\label} \def\eeq{\end{equation}{q1} [a]\,q^b+[b]\,q^{-a}=[a+b] , \eeq we can rewrite (\ref{2.8})-(\ref{2.11}) in the following way: \begin{equation}\label} \def\eeq{\end{equation}{q2} \wh{U}_1 \wh{U}_3 = q^{-1} \wh{U}_3 \wh{U}_1, \;\;\, \wh{U}_2 \wh{U}_4 = q^{-1} \wh{U}_4 \wh{U}_2, \;\;\, \wh{U}_1 \wh{U}_2 = \wh{U}_2 \wh{U}_1, \;\;\, \wh{U}_3 \wh{U}_4 = \wh{U}_4 \wh{U}_3 , \eeq \begin{equation}\label} \def\eeq{\end{equation}{q3} \wh{U}_1 \wh{U}_4 - q^{-1} \wh{U}_4 \wh{U}_1 = q^{-1} {\rm Det}U \, q^{p/\hbar},\;\;\;\;\; \wh{U}_3 \wh{U}_2 - q\wh{U}_2 \wh{U}_3 = - {\rm Det}U \,q^{-p/\hbar}. \eeq The relations (\ref{q3}) are well known in the theory of $q$-oscillators ($q$-bosons).\REF{20} Recall that $q$-analogues of creation, annihilation, and number operators form the deformed Heisenberg algebra defined by the commutation relations \begin{equation}\label} \def\eeq{\end{equation}{q3.1} a\, a^+ - q \, a^+ a = N^{-1},\;\; N\, a = q^{-1}\, a\, N,\;\; N\, a^+ = q\, a^+ N \, , \eeq and they can be realized in terms of multiplication and difference operators: \begin{equation}\label} \def\eeq{\end{equation}{q4} a^+=z,\;\;a=z^{-1} [z\,\partial_z],\;\; N = q^{z\,\partial_z} . \eeq Using two pairs of generators of the deformed Heisenberg algebra, one can construct the generators of $U_q(sl(2))$: $l_+ = a_1^+\,a_2$, $l_- = a_2^+\,a_1$, $q^{l_3}=N_1^{1/2}N_2^{-1/2}$. Applying here the representation (\ref{q4}) one gets \begin{equation}\label} \def\eeq{\end{equation}{q33} l_+ = z_1 z_2^{-1} [z_2\partial_2 ],\;\;\;\;\; l_- = z_2 z_1^{-1} [z_1\partial_1 ],\;\;\;\;\; q^{l_3} = q^{\frac{1}{2} (z_1 \partial_1 - z_2 \partial_2)}. \eeq The Casimir operator (\ref{L8}) of $U_q(sl(2))$ in this realization is given by \begin{equation}\label} \def\eeq{\end{equation}{q4.5} C= q\,N_1 N_2 + q^{-1} N_1^{-1} N_2^{-1}. \eeq Now, comparing, (\ref{q2})-(\ref{q3}) with (\ref{q3.1}), it is easy to conclude that the pairs $(\wh{U}_1,\wh{U}_4)$ and $(\wh{U}_2,\wh{U}_3)$ are similar to two pairs of $q$-boson operators. Taking into account the Weyl-like form of relations (\ref{q2}) and having already found explicit expressions (\ref{2.14})-(\ref{2.15}) for the generators of algebra ${\cal U}_0$, we get an answer for $D$ and $\wh{U}$ in terms of $q$-oscillators. More precisely, a straightforward calculation allows to verify the following statement: \begin{prop}\label{P9} Equations (\ref{q2})-(\ref{q3}) have the family of solutions: \begin{equation}\label} \def\eeq{\end{equation}{q6} q^{p/\hbar} = q\,N_1\,N_2,\;\;\;\; \wh{U} = \left( \begin{array}} \def\er{\end{array}{cc} \alpha} \def\be{\beta} \def\ga{\gamma_0 \,a_1 \, N_1^{\alpha} \def\be{\beta} \def\ga{\gamma}\,N_2^{-\be} &\be_0\, a_2^+\,N_1^{\be}\,N_2^{-\alpha} \def\be{\beta} \def\ga{\gamma} \\ [1mm] -\ga_0\, a_2 \,N_1^{-(1+\be)}\, N_2^{\alpha} \def\be{\beta} \def\ga{\gamma} & \dl_0\,a_1^+\,N_1^{-\alpha} \def\be{\beta} \def\ga{\gamma}\,N_2^{1+\be} \er\right), \eeq where $\;\alpha} \def\be{\beta} \def\ga{\gamma_0\,\dl_0=q\,\be_0\,\ga_0$. \end{prop} Let us note that this form of $\wh{U}$ is consistent with the condition (\ref{2.5}). Taking into account the connection formula (\ref{2.7}) and applying to the generators $a_i$, $a_i^+$, $N_i$ the representation (\ref{q4}), one obtains from (\ref{q6}) a family of representations of the algebra ${\cal U}$. To select some of them, we have to impose an additional condition. As mentioned above (see (\ref{3.05})-(\ref{3.15})), in the non-deformed case substitution of the generating matrix $U_0$ in the formula (\ref{1.7}) gives the matrix $L_0$ which exactly coincides with the limit version of the matrix (\ref{L5}). It is natural to suppose that the generating matrix $U$ corresponding to deformed algebra produces in the same way the matrix (\ref{L5}) itself. Bearing in mind the property (\ref{2.130}), we obtain the following \begin{prop}\label{P10} The condition $\wh{U}D\wh{U}^{-1}=L$, where $L$ is the matrix (\ref{L5}), $D$ is given by \begin{equation}\label} \def\eeq{\end{equation}{q5.5} D = \left(\begin{array}} \def\er{\end{array}{cc} q^{p/\hbar} & \\ & q^{-p/\hbar} \er\right) = \left(\begin{array}} \def\er{\end{array}{cc} q\,N_1\,N_2 & \\ & q^{-1}\,N_1^{-1}\,N_2^{-1} \er\right) \, , \eeq and $\wh{U}$ is given by (\ref{q6}), imposes the following restrictions: \begin{equation}\label} \def\eeq{\end{equation}{q8} \alpha} \def\be{\beta} \def\ga{\gamma+\be+\frac{1}{2}=0,\;\;\; \alpha} \def\be{\beta} \def\ga{\gamma_0=q\,\ga_0 ,\;\;\; \be_0= \dl_0. \eeq \end{prop} Substitution of (\ref{q8}) into (\ref{q6}) completes a description of $\wh{U}$ in terms of $q$-oscillators. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf B. Connection with $\bf (T^*{\cal B})_q$. } \hspace*{6mm} Now we are going to develop another approach to constructing representations of $\cal U$. It is more universal since it is based on the connection (which takes place for arbitrary quantum Lie algebra) of the algebra ${\cal L}$ (see Definition \ref{DL}) with $(T^*\B)_q$ and on the interpretation of the deformed Borel subgroup $\B_q$ as a quantum model space. To clarify the announced connection we start with the following theorem (this is a version of the theorem given in Ref.10 for $L$-operators with nonultralocal relations) \begin{theorem} \label{ABL} Let the matrices $A$ and $B$ obey the relations of type (\ref{1.1}): \begin{equation}\label} \def\eeq{\end{equation}{4.1} R_{\pm }\up{1}{A} \, \up{2}{A} \,=\, \up{2}{A} \, \up{1}{A} R_{\pm} \ , \;\;\;\; R_{\pm } \up{1}{B} \, \up{2}{B} \,=\, \up{2}{B} \, \up{1}{B} R_{\pm} \eeq and the additional exchange relation \begin{equation}\label} \def\eeq{\end{equation}{4.2} \up{1}{A} \, \up{2}{B} \,=\, \up{2}{B} \, \up{1}{A} R_{+} \ , \;\;\;\; \up{2}{A} \, \up{1}{B}R_{-} \,=\, \up{1}{B} \, \up{2}{A} \ . \eeq Then the $L$-operator constructed by means of similarity transformation \begin{equation}\label} \def\eeq{\end{equation}{4.3} L=ABA^{-1} \eeq satisfies the relation (\ref{1.3}). \end{theorem} {\it Remark:} \ } \def\proof{{\it Proof}\ Since (\ref{4.1}) defines a quantum group structure, $A^{-1}$ in (\ref{4.3}) should be understood as an antipode of $A$. \proof of Theorem \ref{ABL} is straightforward : \[ \begin{array}} \def\er{\end{array}{c} \up{1}{L} R^{-1}_{-} \up{2}{L}R_{-}= \up{1}{A} \up{1}{B}(\up{1}{A})^{-1} R_{-}^{-1} \up{2}{A} \up{2}{B} (\up{2}{A})^{-1} R_{-}= \up{1}{A} \up{1}{B} \up{2}{A} R_{-}^{-1} (\up{1}{A})^{-1} \up{2}{B} (\up{2}{A})^{-1}R_{-}=\\ \\ =\up{1}{A}\,\up{2}{A}\,\up{1}{B}\,\up{2}{B}R_{+}^{-1}(\up{1}{A})^{-1} (\up{2}{A})^{-1} R_{-}= R_{+}^{-1} \up{2}{A}\,\up{1}{A}\,\up{2}{B}\,\up{1}{B} R_{-} (\up{2}{A})^{-1} (\up{1}{A})^{-1}= \\ \\ = R_{+}^{-1} \up{2}{A} \up{2}{B}\up{1}{A}R_{+} (\up{2}{A})^{-1} \up{1}{B} (\up{1}{A})^{-1}= R_{+}^{-1}\up{2}{A} \up{2}{B} (\up{2}{A})^{-1} R_{+} \up{1}{A} \up{1}{B}(\up{1}{A})^{-1}= R_{+}^{-1} \up{2}{L} R_{+}\up{1}{L}.\er \] Thus, for a given quantum group $G_q$, the algebra $\cal L$ is embedded into the algebra generated by entries of $A$ and $B$ obeying (\ref{4.1})-(\ref{4.2}). To argue that (\ref{4.1})-(\ref{4.2}) describe a $q$-analogue of $T^*\B$, let us notice that the non-symmetric (with respect to $R$-matrices) form of the relations (\ref{4.2}) imposes some restriction on the structure of the matrices $A$ and $B$. Say, if $R_+$ is an upper triangular matrix, then $A$ and $B$ must be upper and lower triangular, respectively. Therefore, one may think of $A$ and $B$ as coordinates in the deformed Borel subgroup ${\cal B}_q$ and in the dual quantum space, respectively. In other words, the matrices $A$ and $B$ are coordinate and momentum on the deformed phase space $(T^*\B)_q$ respectively. Thus (\ref{4.1})-(\ref{4.2}) may be regarded as a definition of $(T^*\B)_q$ (for additional comments see Ref.10). We should underline here that, although the matrices $A$ and $B$ look similarly on quantum level, they transform into different objects when $q\rightarrow} \def\lar{\leftarrow} \def\ov{\overline 1$. Indeed, in the limit $q\rightarrow} \def\lar{\leftarrow} \def\ov{\overline 1$ one has $L \rightarrow} \def\lar{\leftarrow} \def\ov{\overline I + \ga\hbar L_0$ and the corresponding limit forms of $A$ and $B$ are \begin{equation}\label} \def\eeq{\end{equation}{4.45} A \rightarrow} \def\lar{\leftarrow} \def\ov{\overline A_0, \;\;\; B \rightarrow} \def\lar{\leftarrow} \def\ov{\overline I+\ga\hbar B_0, \eeq where $A_0$ is a group-like element, whereas $B_0$ is rather an element of algebra (see (\ref{3.35b}) as an example of $A_0$, $B_0$ for $sl(2)$). Comparing the statements of Theorems 1 and \ref{ABL} and taking into account the equality (\ref{2.130}), we get the formula \begin{equation}\label} \def\eeq{\end{equation}{4.115} L = A\, B \, A^{-1} = \wh{U} \, D \, \wh{U}^{-1}, \eeq which points out a possibility to construct the matrix $\wh{U}$ obeying (\ref{2.8})-(\ref{2.9}) via the generators of $(T^*\B)_q$. This connection is very important; below we consider it for $SL_q(2)$ in all details. Now let us turn to the example of $SL_q(2)$. For $R_{\pm}$ defined as in (\ref{1.4}) one can choose \begin{equation}\label} \def\eeq{\end{equation}{4.4} A=\left(\begin{array}} \def\er{\end{array}{cc} a & c \\ 0 & a^{-1} \er \right) ,\ \ \ \ B=\left(\begin{array}} \def\er{\end{array}{cc} b & 0 \\ d & b^{-1} \er\right) . \eeq Explicit relations for the generators of $(T^*\B)_q$ following from (\ref{4.1})-(\ref{4.2}) are \begin{equation}\label} \def\eeq{\end{equation}{4.5} a\,c=q^{-1}c\,a,\ \ \ b\,c=q^{1/2}c\,b,\ \ \ a\,b=q^{1/2}b\,a; \eeq \begin{equation}\label} \def\eeq{\end{equation}{4.6} b\,d=q^{-1}d\,b,\ \ \ a\,d=q^{1/2}d\,a,\ \ \ c\,d=q^{-1/2}d\,c+q^{-1/2}\omega\, b^{-1}a. \eeq Performing the following decomposition \begin{equation}\label} \def\eeq{\end{equation}{4.7} d=d_0+d_1=d_0+q^{1/2}c^{-1}b^{-1}a \, , \eeq we transform (\ref{4.6}) to homogeneous form: \begin{equation}\label} \def\eeq{\end{equation}{4.8} b\,d_0=q^{-1}d_0\, b,\ \ \ \ a\, d_0=q^{1/2}d_0 \, a, \ \ \ \ c\, d_0=q^{-1/2}d_0 \, c . \eeq Thus, (\ref{4.5}) and (\ref{4.8}) describe four variables obeying Weyl-like commutation relations. Using the jargon of conformal field theory, we shall call these formulae "free field representation" and the generators $a$, $b$, $c$, $d_0$ "free field" variables. {\it Remark:} \ } \def\proof{{\it Proof}\ The last of equations (\ref{4.6}) is nothing but a commutation relation entering the definition of deformed Heisenberg algebra. Indeed, comparing (\ref{4.5})-(\ref{4.6}) with (\ref{q3.1}), one can establish the following correspondence ($\rho$ stands for arbitrary numerical constant): $$ c \sim N^{\rho}\,a^{+},\;\;\; d \sim -\omega \, N^{-1/2-\rho}\,a,\;\;\; b^{-1}\,a \sim q^{\rho}\,N^{-3/2}. $$ Thus, the transformation (\ref{4.7}) can be interpreted as "bosonization" of $q$-oscillators. Now, substituting (\ref{4.4}) in (\ref{4.3}), we get \begin{equation}\label} \def\eeq{\end{equation}{4.9} \begin{array}} \def\er{\end{array}{c} L = q^{1/2} \left( \begin{array}} \def\er{\end{array}{cc} a & c \\ 0 & a^{-1} \er \right) \left( \begin{array}} \def\er{\end{array}{cc} b & 0 \\ d & b^{-1} \er \right) \left( \begin{array}} \def\er{\end{array}{cc} a^{-1} & -q\,c \\ 0 & a \er \right) = \\ \\ = \left( \begin{array}} \def\er{\end{array}{cc} q\, (b+b^{-1}) + a^{-1}c\, d_0 & - q^2 a\, c\, (b+q \, a^{-1} c\, d_0) \\ (a\,c)^{-1} (b^{-1} + q^{-1} a^{-1}c\, d_0) & -q^2 a^{-1} c\, d_0 \er \right). \er \eeq This matrix provides a "free field" realization of the algebra $\cal L$ for $U_q(sl(2))$. Note that the additional scaling factor $q^{1/2}$ was introduced in (\ref{4.9}) to ensure a coincidence of the Casimir operators calculated by formulae (\ref{1.6}) for the matrix (\ref{4.9}): \begin{equation}\label} \def\eeq{\end{equation}{4.10} K_1 = q^2 (b+b^{-1}), \;\;\; K_2 = q^3 \eeq with those for the matrix (\ref{L5}). In fact, we redefined the matrix $B$ in (\ref{4.4}) as \begin{equation}\label} \def\eeq{\end{equation}{4.95} \wt{B} = q^{1/2} B . \eeq Comparing the Casimir operator $K_1$ given by (\ref{4.10}) with one given by (\ref{L6}), we identify the operator $b$ with the power of the operator of spin $\wh{j}$ : \begin{equation}\label} \def\eeq{\end{equation}{4.105} b = q^{2 \wh{j} +1}. \eeq It follows from (\ref{4.10}) that matrix $L$ contains only three independent variables (it is easy to see from the explicit form (\ref{4.9}) that these are $b$, $ac$ and $a^{-1}cd_0$). Moreover, direct calculation using (\ref{4.5}),(\ref{4.8}) shows that all elements of the matrix $L$ commute with operator $b$. That agrees with the property (\ref{1.12}). Now exploiting the connection described by formula (\ref{4.115}), one can obtain an exact expression for $\wh{U}$. \begin{theorem} \label{main} The algebra ${\cal U}\equiv \{\wh{U},p\}$ with defining relations (\ref{2.8})-(\ref{2.9}) has the following realization in terms of generators $a$, $b$, $c$, $d_0$ \begin{equation}\label} \def\eeq{\end{equation}{5.00} b = q^{p/\hbar},\;\;\;\; \wh{U} = \left( \begin{array}} \def\er{\end{array}{cc} \frac{1}{\omega}\, a\,(b+a^{-1}\,c\,d_0)\, e^{-\frac{i \xi}{2}} & c\,e^{\frac{i \xi}{2}} \\ [1mm] \frac{1}{\omega}\, c^{-1}(b^{-1} + q^{-1} a^{-1}\,c\,d_0)\, e^{-\frac{i\xi}{2}} & a^{-1}\, e^{\frac{i \xi}{2}} \er \right), \eeq where $\omega\equiv q-q^{-1}$,\ $d_0$ is defined in (\ref{4.7}), and \begin{equation}\label} \def\eeq{\end{equation}{5.15} e^{i \xi}= a^{-1}\;b^{\ga}\;c^{-1}\;d_0^{-1} \eeq with $\ga$ being an arbitrary constant. \end{theorem} This theorem gives a "free field" representation of the algebra $\cal U$. Let us remark that the remaining freedom in (\ref{5.15}) corresponds only to canonical transformations (since $\xi$ and $p$ are conjugate variables). The formulated theorem will be proved in several steps. First, we introduce a lower-triangular matrix which diagonalizes the matrix $\wt{B}$: \begin{equation}\label} \def\eeq{\end{equation}{5.1} V = \left( \begin{array}} \def\er{\end{array}{cc} v_1 & 0 \\ v_3 & v_2 \er \right), \;\;\; \wt{B} = V \wt{B}_0 V^{-1}, \;\;\; \wt{B}_0 = \left(\begin{array}} \def\er{\end{array}{cc} q^{1/2} b & 0 \\ 0 & q^{1/2} b^{-1} \er \right) \equiv q^{1/2} B_0 . \eeq \begin{prop} A possible solution for the matrix $V$ is \begin{equation}\label} \def\eeq{\end{equation}{5.3} v_1 =v_1(b),\;\;\; v_2 =v_2(b),\;\;\; v_3 = d \,v_1(b)\,f(b), \eeq where $v_1(b)$, $v_2(b)$ are arbitrary functions of $b$ and $f(b) = (b-qb^{-1})^{-1}$. \end{prop} Thus, matrix $L$ given by (\ref{4.9}) admits a decomposition of the form: \begin{equation}\label} \def\eeq{\end{equation}{5.5} L=\wh{U}_0\, \wt{B}_0\, \wh{U}_0^{-1}, \;\;\;\; \wh{U}_0 = A\,V. \eeq However, this diagonalization is not unique. Using an arbitrary power of the diagonal matrix $Q$, which depends on the variable conjugate to $b$, \begin{equation}\label} \def\eeq{\end{equation}{5.6} Q = \left( \begin{array}} \def\er{\end{array}{cc} e^{i\xi} & \\ & e^{-i\xi} \er \right), \;\;\;\; b\,e^{i\xi} = q\,e^{i\xi}\,b, \eeq we obtain a family of diagonalizing matrices: \begin{equation}\label} \def\eeq{\end{equation}{5.7} L= \wh{U}_{\dl}\,\wt{B}_{\dl}\,\wh{U}_{\dl}^{-1}, \;\;\; \wh{U}_{\dl}= A\,V\,Q^{\dl}, \;\;\; \wt{B}_{\dl} = Q^{-\dl} \wt{B}_0 \, Q^{\dl} = q^{\dl}\,\wt{B}_0 = q^{\dl+1/2}\, B_0. \eeq An explicit form of the diagonalizing matrix is \begin{equation}\label} \def\eeq{\end{equation}{5.8} \wh{U}_{\dl} = A\,V\,Q^{\dl} = \left( \begin{array}} \def\er{\end{array}{cc} (a\,v_1 +c\,d\,v_1 f)\, e^{i\dl \xi} & c\,v_2 \,e^{-i\dl \xi} \\ a^{-1} d\,v_1 f\, e^{i\dl \xi} & a^{-1}v_2 \, e^{-i\dl \xi} \er \right). \eeq Here we should describe a new object $e^{i \xi}$ which appeared in the matrix $\wh{U}$. We assume that the following Weyl-like relations hold: \begin{equation}\label} \def\eeq{\end{equation}{5.10} a\, e^{i \xi}= q^{\alpha} \def\be{\beta} \def\ga{\gamma} e^{i \xi}\, a, \;\;\; b\, e^{i \xi}= q\,e^{i \xi}\, b, \;\;\; c\, e^{i \xi}= q^{\be} e^{i \xi}\, c,\;\;\; d_0\, e^{i \xi}= q^{\ga} e^{i \xi}\,d_0. \eeq \begin{prop} The set of equations (\ref{5.10}) is equivalent to \begin{equation}\label} \def\eeq{\end{equation}{5.11} e^{i\xi}= a^{\be+(\ga-1)/2}\;b^{\ga}\;c^{(\ga-1)/2-\alpha} \def\be{\beta} \def\ga{\gamma}\;d_0^{-1}. \eeq \end{prop} Now we have to remind that the matrix $U$ (and $\wh{U}$ as well) described in the Theorem 1 has to satisfy the relation (\ref{1.8}) or, equivalently, to the relation \begin{equation}\label} \def\eeq{\end{equation}{5.85} \up{1}{B_0} \,\up{2}{\wh{U}_{\dl}} = \up{2}{\wh{U}_{\dl}} \, \up{1}{B_0} \, \sigma, \eeq where $\sigma$ and $B_0$ were introduced in (\ref{1.8}) and (\ref{5.1}), respectively. A straightforward calculation using (\ref{4.5})-(\ref{4.6}) leads to the following. \begin{prop} The matrix $\wh{U}_{\dl}$ given by (\ref{5.8}) satisfies the relation (\ref{5.85}) only for $ \dl = -1/2$. \end{prop} It is worth mentioning that such a choice of $\dl$ exactly compensates the renormalization of the matrix $B$ in (\ref{4.95}), i.e., \ $\wt{B}_{-1/2} = B_0$. Bearing in mind the formula (\ref{4.7}), one can rewrite (\ref{5.8}) for $\dl = \begin{array}} \def\er{\end{array}{c}-\frac{1}{2} \er$ as follows \begin{equation}\label} \def\eeq{\end{equation}{5.9} \wh{U} \equiv \wh{U}_{-1/2} = \left( \begin{array}} \def\er{\end{array}{cc} a(b+a^{-1}\,c\,d_0)\,w\, e^{-\frac{i \xi}{2}} & c\,v\,e^{\frac{i \xi}{2}} \\ c^{-1}(b^{-1} + q^{-1} a^{-1}\,c\,d_0)\,w\,e^{-\frac{i \xi}{2}} & a^{-1}\,v \, e^{\frac{i \xi}{2}} \er \right), \eeq where $w \equiv f(b) v_1(b)$,\ $v \equiv v_2(b)$. Finally, a direct check shows (see the Appendix C) that the matrix (\ref{5.9}) obeys eqs.$\!$ (\ref{2.8})-(\ref{2.9}) if the functions $w$, $v$ are constant (we chose them as follows: $v(b)=1$, $w(b)=\frac{1}{\omega}$) and the coefficients in (\ref{5.10})-(\ref{5.11}) satisfy the conditions $\be \!=\! -\alpha} \def\be{\beta} \def\ga{\gamma$, \ $\ga \!=\! \alpha} \def\be{\beta} \def\ga{\gamma-\be-1 \!=\! 2\alpha} \def\be{\beta} \def\ga{\gamma-1$. Thus, Theorem \ref{main} is proven. Let us end the discussion of relation of $(T^*\B)_q$ to algebras $\cal L$ and $\cal U$ with one more statement: \begin{theorem} The algebra generated by coordinates on $(T^*\B)_q$ is isomorphic to the algebra generated by entries of the matrix $L$ and $Q$. \end{theorem} \proof. Formulae (\ref{4.9}) and (\ref{5.15}) provide explicit expressions for entries of $L$ and $Q$ via the generators $a$, $b$, $c$, $d_0$ (up to unessential canonical transformation in (\ref{5.15})). Conversely, suppose matrix $L$ and the element $e^{i\xi}$ are given. Then, as it follows from (\ref{4.9}), one can construct from entries of $L$ the combinations $b$, $ac$ and $a^{-1}cd_0$. Together with (\ref{5.15}) this allows to recover the ``coordinates'' $a$, $b$, $c$, $d_0$. Although we considered this theorem only for the case of $SL_q(2)$, there is an evidence that it holds for the generic case. Indeed, in the case of $G_q=SL_q(N)$ a point on the quantum bundle $(T^*\B)_q$ is parameterized by $N\times N$ matrices $A$ and $B$. As above, the matrix $L=ABA^{-1}$ satisfies (\ref{1.3}) and therefore its entries generate the corresponding algebra $\cal L$. However, the dimension of $(T^*\B)_q$ exceeds the dimension of $\cal L$ : \ ${\rm dim}\, (T^*\B)_q - {\rm dim}\, {\cal L} = (N^2+N-2)-(N^2-1)=N-1$. It us very probable that the remaining $(N-1)$ generators are exactly those that enter the diagonal unimodular $N\times N$ matrix $Q$. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf C. Explicit representation.} \hspace*{6mm} Now we face the problem of constructing of an explicit representation for the generators $a$, $b$, $c$, $d_0$. A Weyl-like form of the commutation relations (\ref{4.5}),(\ref{4.8}) points out the possibility of getting a realization for these generators in terms of two pairs of canonical variables. This also means (due to the interpretation of (\ref{4.7}) as "bosonization" of $q$-oscillators) that the generators $a$, $b$, $c$, $d$ admit a realization via $q$-oscillators. Evidently, such a representation is not unique. It is natural to realize $a$, $b$, $c$, $d_0$ as operators acting on the $q$-analogue of the space $D(z_1,z_2)$. We shall denote this space as $D_q(z_1,z_2)$. The space $D_q(z_1,z_2)$ is spanned on the basic vectors of form (remember that $[x]$ stands for $q$-numbers) \begin{equation}\label} \def\eeq{\end{equation}{4.01} |j,m\rangle =\frac{z_1^{j+m}z_2^{j-m}}{\sqrt{[j+m]![j-m]!}},\;\;\;\;j= \begin{array}} \def\er{\end{array}{lllll}0,&\!\!\frac{1}{2},&\!\!1,&\!\!\frac{3}{2},&\!\! ...\er \;\;\;\; m=-j,..,j. \eeq One can define on $D_q(z_1,z_2)$ such a scalar product that the system (\ref{4.01}) is orthonormal, that is $ <j,m|j^{\prime},m^{\prime}>=\delta_{j j^{\prime}} \delta_{m m^{\prime}}$. {\it Remark:} \ } \def\proof{{\it Proof}\ This scalar product is a deformation of (\ref{3.3}). Its explicit form makes use of the $q$-exponent and the Jackson integral. See Ref.20 for details. In all formulae concerning the space $D_q(z_1,z_2)$ we suppose that $q$ is chosen as described in Sec.$\!$ II (i.e., it belongs either to the real axis or to the unit circle at the complex plane). In this case an analogue of the rule of conjugation (\ref{3.5}) is \begin{equation}\label} \def\eeq{\end{equation}{4.03} (z_i)^{*}=z_i^{-1}\;[z_i\partial_i],\;\;\;\;\; (z_i\partial_i)^{*}=z_i\partial_i. \eeq The formulae (\ref{4.105}) and (\ref{5.00}) imply that the generator $b$ is a power of the operator of spin. Hence, on the space $D_q(z_1,z_2)$ it is given by \begin{equation}\label} \def\eeq{\end{equation}{5.16} b = q^{z_1\,\partial_1 + z_2\,\partial_2 + 1 } = q\,N_1\,N_2. \eeq Next, let us remind that we already know the limit versions of the generators $a$, $b$, $c$, $d$ (see Proposition \ref{P6}; one should take into account the rescaling (\ref{4.95})). Their appropriate deformations for generic $q$ are described by \begin{prop}\label{P15} The set of operators (with arbitrary constants $\lambda} \def\up{\stackrel} \def\wt{\widetilde_i$, $\nu_i$) \begin{equation}\label} \def\eeq{\end{equation}{5.17} \begin{array}} \def\er{\end{array}{ll} a=q^{\lambda} \def\up{\stackrel} \def\wt{\widetilde_0}\,z_1^{-1/2}\,N_1^{\lambda} \def\up{\stackrel} \def\wt{\widetilde_1}, & c=q^{\nu_0}\,z_1^{-1/2}\,z_2\, N_1^{\lambda} \def\up{\stackrel} \def\wt{\widetilde_1-2}\,N_2^{\nu_2}, \\ \\ b=q\,N_1\,N_2, & d=-q^{\lambda} \def\up{\stackrel} \def\wt{\widetilde_0-\nu_0+\nu_2}\,z_2^{-1}\,(N_2-N_2^{-1})\,N_1\,N_2^{-\nu_2} \er \eeq satisfies (\ref{4.5})-(\ref{4.6}) and gives in the limit $\ga\rightarrow} \def\lar{\leftarrow} \def\ov{\overline 0$ the generators found in (\ref{3.35b}). \end{prop} Although due to Theorem \ref{main} this proposition gives a family of representations for $\cal U$, we again should impose an additional condition using the matrix (\ref{L5}) as a standard (justification for this trick was given above). \begin{prop}\label{P16} Matrix $L$ given by (\ref{4.9}) coincides with the matrix (\ref{L5}) taken in the representation (\ref{q33}) provided that \begin{equation}\label} \def\eeq{\end{equation}{5.19} b = q\,N_1\,N_2,\;\;\;a\,c= q^{-1/2}\,z_1^{-1}\,z_2\,N_1^{-1/2} \,N_2^{-1/2},\;\;\; a^{-1}\,c\,d_0=-N_1^{-1}\,N_2. \eeq \end{prop} Comparing the statements of Propositions \ref{P15} and \ref{P16}, we derive: \begin{equation}\label} \def\eeq{\end{equation}{5.20} \begin{array}} \def\er{\end{array}{ll} a=q^{\lambda} \def\up{\stackrel} \def\wt{\widetilde_0}\,z_1^{-1/2}\,N_1^{3/4},\ & b=q\,N_1\,N_2, \\ c=q^{\nu_0}\, z_1^{-1/2}\,z_2\,N_1^{-5/4}\,N_2^{-1/2}, & \\ d_0=-q^{\lambda} \def\up{\stackrel} \def\wt{\widetilde_0-\nu_0-1/2}\, z_2^{-1}\,N_1\,N_2^{3/2},\ & q^{\lambda} \def\up{\stackrel} \def\wt{\widetilde_0+\nu_0}= q^{-1/8} . \er \eeq Substituting (\ref{5.20}) into (\ref{5.15}) (and remember that (\ref{5.15}) is defined only up to a coefficient), we get \begin{equation}\label} \def\eeq{\end{equation}{5.21} e^{i\xi} = q^{2\epsilon}\,z_1\,N_1^{\ga-1/2}\,N_2^{\ga-1}, \eeq where $\ga$ and $\epsilon$ are arbitrary. Finally, substituting (\ref{5.20})-(\ref{5.21}) into (\ref{5.9}), we obtain (one should remember $U$ and $\wh{U}$ are defined only up to arbitrary scaling factor) \begin{equation}\label} \def\eeq{\end{equation}{5.22} \wh{U} = \! \left( \begin{array}} \def\er{\end{array}{cc} \frac{1}{\omega}\,\alpha} \def\be{\beta} \def\ga{\gamma_0\,z_1^{-1}\, N_1^{1-\ga/2}\,N_2^{3/2-\ga/2} (N_1-N_1^{-1}) & \be_0 \,z_2\,N_1^{\ga/2-3/2} N_2^{\ga/2-1} \\ & \\ - \frac{1}{\omega}\, q^{-1} \alpha} \def\be{\beta} \def\ga{\gamma_0\, z_2^{-1} N_1^{1/2-\ga/2}\, N_2^{1-\ga/2} (N_2-N_2^{-1}) & \be_0\,z_1\,N_1^{\ga/2-1} N_2^{\ga/2-1/2} \er \right). \eeq It is easy to check that the family of matrices (\ref{5.22}) exactly coincides with what was obtained in $q$-oscillator approach (see Propositions \ref{P9} and \ref{P10}). \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf D. Quantum Clebsch-Gordan coefficients. } \hspace*{6mm} Using the connection formula (\ref{2.7}) we get from (\ref{5.22}) a family of matrices $U$ which provide possible representations of the algebra $\cal U$. It is natural to study an action of the entries of these matrices on the space $D_q(z_1,z_2)$ described above. On the basic vectors (\ref{4.01}) these operators act as follows: \begin{equation}\label} \def\eeq{\end{equation}{6.1} \begin{array}} \def\er{\end{array}{c} U_1\, |j,m\rangle = C_1\,q^{\frac{1}{2}(j-m+1)} \sqrt{\frac{[j+m]}{[2j+1]}}\, |j-\frac{1}{2},m-\frac{1}{2}\rangle , \\ \\ U_2\, |j,m\rangle = C_2\,q^{-\frac{1}{2}(j+m)} \sqrt{\frac{[j-m+1]}{[2j+1}]}\, |j+\frac{1}{2},m-\frac{1}{2}\rangle , \\ \\ U_3\, |j,m\rangle =-C_3\,q^{-\frac{1}{2}(j+m+1)} \sqrt{\frac{[j-m]}{[2j+1]}}\, |j-\frac{1}{2},m+\frac{1}{2}\rangle , \\ \\ U_4\, |j,m\rangle =C_4\,q^{\frac{1}{2}(j-m)} \sqrt{\frac{[j+m+1]}{[2j+1]}}\, |j+\frac{1}{2},m+\frac{1}{2}\rangle , \er \eeq where the coefficients $C_i$ do not depend on $m$. Note that, similarly to the classical case, the operators $U_i$ correspond to the basic shifts on the model space. Comparing the matrix elements $<j^{\prime},m^{\prime}|U_i|j,m>$ following from (\ref{6.1}) with values of CGC for $U_q(sl(2))$ given by $q$-analogue of the Van-der-Waerden formula,\REF{18,21} which for the decomposition of $V_j\otimes V_{1/2}$ looks like following \[ \left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime } \\ m & m^{\prime } & m^{\prime \prime } \er \right\}_q = \delta_{m^{\prime \prime},m + m^{\prime}}\; \left(\frac{[j+\frac 12-j^{\prime \prime }]![j+j^{\prime \prime }-\frac 12]![j^{\prime \prime }+\frac 12-j]!}{[j+j^{\prime \prime }+\frac 32]!} \right)^{1/2} \times \] \begin{equation}\label} \def\eeq{\end{equation}{6.2} \times q^{\frac{1}{2} (j+\frac{1}{2} -j^{\prime \prime})(j+ j^{\prime \prime} +\frac{3}{2}) + jm^{\prime} -\frac{1}{2} m} \times \eeq \[ \times \sum\limits_{r\geq 0}\frac{(-1)^r q^{-r(j+j^{\prime \prime}+ \frac 32)} \left( [j+m]![j-m]![j^{\prime \prime}+m^{\prime\prime }]![j^{\prime \prime }-m^{\prime \prime }]![2j^{\prime \prime }+1]\right) ^{1/2} } {[r]![j+\frac 12-j^{\prime \prime }-r]![j-m-r]![\frac 12+m^{\prime } -r]![j^{\prime\prime }-\frac 12+m+r]![j^{\prime \prime }-j-m^{\prime }+r]!},\] we establish the following correspondence \[ \! \begin{array}} \def\er{\end{array}{ll} \langle j^{\prime \prime} , m^{\prime \prime} \mid U_1\mid j, m\rangle= \delta_{j^{\prime \prime },j- \frac 12 }\, \alpha} \def\be{\beta} \def\ga{\gamma_0\,q^{(1-\ga/2)j-1/2}\, \left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime} \\ m & -\frac 12 & m^{\prime \prime} \er \right\}_q , \\ \langle j^{\prime \prime} , m^{\prime \prime} \mid U_2 \mid j, m\rangle= \delta_{j^{\prime \prime },j+ \frac 12 }\, \be_0\,q^{(\ga/2-1)j}\,\left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime} \\ m & -\frac 12 & m^{\prime \prime} \er \right\}_q , \\ \langle j^{\prime \prime} , m^{\prime \prime} \mid U_3 \mid j, m\rangle= \delta_{j^{\prime \prime },j- \frac 12 }\, \alpha} \def\be{\beta} \def\ga{\gamma_0\,q^{(1-\ga/2)j-1/2}\, \left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime} \\ m & \frac 12 & m^{\prime \prime} \er \right\}_q, \\ \langle j^{\prime \prime} , m^{\prime \prime} \mid U_4 \mid j, m\rangle= \delta_{j^{\prime \prime },j+ \frac 12 }\, \be_0\,q^{(\ga/2-1)j}\,\left\{\begin{array}} \def\er{\end{array}{ccc} j & \frac 12 & j^{\prime \prime} \\ m & \frac 12 & m^{\prime \prime}\er \right\}_q . \er \] Thus we derive an analogue of Proposition \ref{P7}: \begin{prop}\label{P17} The generators $U_i$ of the algebra $\cal U$ are operators of the basic shifts on the model space for $U_q(sl(2))$ and they generate the $q$-Clebsch-Gordan coefficients corresponding to decomposition of the product $V_j\otimes V_{\frac 12}$ of irreps of $U_q(sl(2))$. \end{prop} {\it Remark:} \ } \def\proof{{\it Proof}\ Putting $\alpha} \def\be{\beta} \def\ga{\gamma_0=q^{1/2}$, $\be_0=1$ and $\ga=2$ in (\ref{5.22}), we get the following generating matrix \begin{equation}\label} \def\eeq{\end{equation}{6.3} U = \left( \begin{array}} \def\er{\end{array}{cc} z_1^{-1}\, [z_1 \partial_1 ] q^{\frac 12 (z_2 \partial_2 +1)} & z_2 \, q^{-\frac 12 z_1 \partial_1 } \\ \\-z_2^{-1}\, [z_2 \partial_2 ] q^{-\frac 12 (z_1 \partial_1 +1)} & z_1 \, q^{\frac 12 z_2 \partial_2 } \er \right) \,\frac{1}{\sqrt{[p/\hbar]} \def\pp{[p/\hbar+1]} \def\pu{[p/\hbar-1]}} \ , \;\;\;\; {\rm Det} U = q^{1/2} \ , \eeq which may be called ``exact'' as it satisfies (\ref{6.1}) with $C_i=1$. The question about unitarity of the matrix (\ref{6.3}) is discussed in Appendix B. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf E. Generalized Wigner-Eckart theorem.} \hspace*{6mm} As we demonstrated in the previous section, entries of the matrix $U_0$ are tensor operators of spin $1/2$ for ${\cal J}=sl(2)$, hence they provide a realization of the Wigner-Eckart theorem. Let us now consider the matrix $U$ from this point of view. The theory of tensor operators for quantum algebras was discussed by many authors (see, e.g., Ref.22). In particular, generalized Wigner-Eckart theorem (in the case of ${\cal J}} \def\Uc{{\cal U}_q=U_q(sl(2))$ reads as follows. \begin{theorem} \label{TWE} Let $l_{+}$, $l_{-}$ and $l_{3}$ be the generators of $U_q(sl(2))$ and let $T^{j\;}_{\;m}$, $m=-j,..,j$ be a set of operators acting on the deformed model space $\cal M$ and obeying the commutation relations \begin{equation}\label} \def\eeq{\end{equation}{WE1} {} [l_3,T^{j\;}_{\;m}]=m\,T^{j\;}_{\;m},\;\;\; {} l_{\pm}\,T^{j\;}_{\;m}\,q^{l_3} - q^{l_3\mp 1}\, T^{j\;}_{\;m}\,l_{\pm} =\sqrt{[j\mp m][j\pm m+1]}\,T^{j\;}_{\;m\pm 1}. \eeq Then the matrix elements of $T^{j\;}_{\;m}$ on $\cal M$ are proportional to $q$-Clebsch-Gordan coefficients: \[ <j^{\prime\prime} m^{\prime\prime} |T^{j\;}_{\;m}|j^{\prime} m^{\prime} > = C_{j^{\prime\prime}j^{\prime}}^{j}\;\left\{\begin{array}} \def\er{\end{array}{ccc}j^{\prime} & j & j^{\prime\prime}\\ m^{\prime} & m & m^{\prime\prime} \er \right\}_q, \] where the coefficients $C_{j^{\prime\prime}j^{\prime}}^j$ do not depend on $m$, $m^{\prime}$, $m^{\prime\prime}$. \end{theorem} Proposition \ref{P17} implies that $U_i$ may be regarded as q-tensor operators. Indeed, using (\ref{5.22}) and (\ref{q33}), one can check that $U_i$ satisfy (\ref{WE1}) (one obtains for $U_i$ deformations of relations (\ref{we20})). Similarly to the classical case we have the following. \begin{prop}\label{P18} The generators $U_i$ of the algebra ${\cal U}$ form a basis for $q$-tensor operators of spin $1/2$, that is components $T^{1/2}_{1/2}$ and $T^{1/2}_{-1/2}$ of any $q$-tensor operator of spin 1/2 can be realized as linear combinations of $U_i$: \begin{equation}\label} \def\eeq{\end{equation}{WE5} T^{1/2}_{-1/2}= \mu(p) \;U_1+\nu(p) \;U_2,\;\;\; T^{1/2}_{1/2}=\mu(p) \;U_3+\nu(p) \;U_4, \eeq where $\mu(p)$ and $\nu(p)$ are functions only of $p$. \end{prop} {\it Remark:} \ } \def\proof{{\it Proof}\ Unlike the classical case, solution (\ref{5.22}) gives a family of matrices $U$. However, the corresponding matrix elements $<j^{\prime\prime} m^{\prime\prime} |U_i|j^{\prime} m^{\prime} > $ differ only by factors which do not depend on $m^{\prime}$, $m^{\prime\prime}$. Thus, any representative of obtained family of matrices $U$ may be used in Proposition \ref{P18}. Let us end the description of the algebra $\cal U$ from the point of view of theory of $q$-tensor operators with the following statement: \begin{prop}\label{P19} The matrices $U$ and $L$ defined in the Theorem 1 obey the relation \begin{equation}\label} \def\eeq{\end{equation}{WE10} R_- \,\up{1}{U}\,\up{2}{L} = \up{2}{L}\,R_+ \,\up{1}{U}. \eeq \end{prop} The proof is straightforward \[ \begin{array}} \def\er{\end{array}{c} R_- \,\up{1}{U}\,\up{2}{L} \, = R_- \,\up{1}{U}\,\up{2}{U}\,\up{2}{D}\,(\up{2}{U})^{-1}= \, \up{2}{U}\, \up{1}{U}\, {\cal R}_- \,\up{2}{D}\,(\up{2}{U})^{-1} = \\ [1mm] =\, \up{2}{U}\, \up{1}{U}\,\up{2}{D}\,\sigma\, {\cal R}_+ \,(\up{2}{U})^{-1} = \, \up{2}{U}\,\up{2}{D}\,\up{1}{U}\, {\cal R}_+ \,(\up{2}{U})^{-1} = \up{2}{U}\,\up{2}{D}\,(\up{2}{U})^{-1}\, R_+ \,\up{1}{U} = \, \up{2}{L}\,R_+ \,\up{1}{U}; \er \] it makes use the relations (\ref{1.8})-(\ref{1.9}) and the property (\ref{RD}). A remarkable fact is that (\ref{WE10}) may be used for definition of $q$-tensor operators instead of (\ref{WE1}). Indeed, in the limit $\ga\rightarrow} \def\lar{\leftarrow} \def\ov{\overline 0$ it turns into \begin{equation}\label} \def\eeq{\end{equation}{WE3} [\up{1}{U_0}, \up{2}{L}_0 ] = \Lambda\, \up{1}{U_0},\;\;\;\; \Lambda=\left(\begin{array}} \def\er{\end{array}{cccc} 1/2& & & \\ &-1/2&1& \\ &1&-1/2& \\ & & &1/2 \er\right). \eeq Using the explicit form of $L_0$ given in (\ref{3.05}), one can easily check that this matrix relation is equivalent to (\ref{we20}). More on $R$-matrix description of $q$-tensor operators is given in Ref.23. \subsection*{ \hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf CONCLUSION.} In this paper we have constructed the $q$-analogue of the phase space $T^*\B$ and clarified its role in description of the model representation of the corresponding quantum group $G_q$. We unraveled a connection between the algebras generating by entries of matrix $(A,B)$, $(U,D)$ and $(L,Q)$. The general formulae were concretized by the example of $G=SL(2)$. An extension of the described scheme to the case of arbitrary group $G$ will definitely improve understanding of the role played by the matrix ${\cal R}(p)$ which so far has been discussed in the literature much less then the standard matrix $R$. The results of this paper can be generalized in several directions even for the case of $SL(2)$. The first is consideration of the matrix $U$ with an auxiliary space corresponding to the higher spin representation. It must lead to an exact form of generating matrix for all CGC. The work in this direction is in progress now. The second point to be discussed is the case of $q$ being a root of unity. The structure of ${\cal R}(p)$ allows to hope that reduction on so-called "good" representations will be quite natural in our formalism. However, this case is to be examined more carefully. \subsection*{\normalsize} \def\ds{\displaystyle\bf \hspace*{6mm} Acknowledgments. } We are grateful to A.Yu. Alekseev, P.P. Kulish and V. Schomerus for stimulating discussions and useful comments. We would like to thank Prof. A. Niemi for hospitality at TFT, University of Helsinki, where this work was begun. This work was partially supported by ISF grant R2H000 and by INTAS grant. \subsection*{ \hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf Appendix A: Proof of Theorem 1. } Using (\ref{1.8})-(\ref{1.9}) together with the identity (\ref{RD}) and taking into account that matrices $\up{1}{D}$, $\up{2}{D}$ and $\sigma$ mutually commute, we check \[ \begin{array}} \def\er{\end{array}{c} \up{1}{L} R_-^{-1} \up{2}{L} R_- = \,\up{1}{U} \up{1}{D} (\up{1}{U})^{-1} R_-^{-1}\up{2}{U}\, \up{2}{D} (\up{2}{U})^{-1} R_- =\\ \\ =\,\up{1}{U} \up{1}{D}\,\up{2}{U} {\cal R}_-^{-1}(p)\, (\up{1}{U})^{-1}\! \up{2}{D} (\up{2}{U})^{-1} R_- = \, \up{1}{U}\, \up{2}{U} \up{1}{D} \sigma\, {\cal R}_-^{-1}(p) \up{2}{D} \sigma\, (\up{1}{U})^{-1} (\up{2}{U})^{-1} R_- =\\ \\ =R_+^{-1} \up{2}{U}\, \up{1}{U} {\cal R}_+ (p) \up{1}{D}\sigma\, {\cal R}_-^{-1}(p) \up{2}{D} \sigma\, {\cal R}_- (p)\, (\up{2}{U})^{-1} (\up{1}{U})^{-1} =\\ \\ =R_+^{-1}\up{2}{U}\,\up{1}{U} \up{2}{D} \sigma\up{1}{D} {\cal R}_- (p) \, (\up{2}{U})^{-1} (\up{1}{U})^{-1} = R_+^{-1} \up{2}{U} \up{2}{D}\, \up{1}{U} {\cal R}_+ (p)\, \sigma\up{1}{D}(\up{2}{U})^{-1} (\up{1}{U})^{-1} =\\ \\ =R_+^{-1} \up{2}{U}\up{2}{D}\, \up{1}{U} {\cal R}_+ (p)\, (\up{2}{U})^{-1}\! \up{1}{D}(\up{1}{U})^{-1} = R_+^{-1} \up{2}{U} \up{2}{D} (\up{2}{U})^{-1} R_+\up{1}{U} \up{1}{D} (\up{1}{U})^{-1} = R_+^{-1} \up{2}{L} R_+ \up{1}{L}. \er\] \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf Appendix B: On conjugation of $\bf U_0$ and $\bf U$. } {}First we consider the matrix $U_0$. Using the rules of conjugation (\ref{3.5}) (and taking into account that $p^*=p$), one can check that the matrix conjugated to $U_0$ does not coincide with $U_0^{-1}$; that is, the matrix $U_0$ itself is not unitary. However, it turns out that the transposed matrix (one should remember that in general $(U^T)^{-1}\neq (U^{-1})^T$ for matrices with non-commuting entries) \[U_0^T = \left( \begin{array}} \def\er{\end{array}{cc} \partial_1 & -\partial_2\\ z_2 & z_1 \er\right)\,\sqrt{\frac{\hbar}{p}} \] satisfies the unitarity condition: \[ (U_0^T)^* = \sqrt{\frac{\hbar}{p}}\,\left(\begin{array}} \def\er{\end{array}{cc} z_1 & \partial_2 \\ -z_2 & \partial_1 \er \right) = (U_0^T)^{-1}. \] In the deformed case (recall that $q$ can be either real or $|q|=1$) the matrix $U$ includes the operator $N$ which conjugates in different ways for the different choices of $q$. Let us consider the matrix $U$ given by (\ref{6.3}). The conjugated matrix can be constructed according to the rules (\ref{4.03}). Using the formula (\ref{q1}), one can check that the unitarity condition $(U^T)^{*}\,U^T=\,U^T\,(U^T)^{*}= I $ (i.e., the same as in the non-deformed case) for the transposed matrix holds only for real $q$. For $|q|=1$ see Ref.13. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf Appendix C: Proof of Theorem 4. } Here we complete the proof of Theorem \ref{main}, i.e., we have to prove that matrix (\ref{5.9}) satisfies (\ref{2.8})-(\ref{2.11}) if the following conditions \hspace*{2cm}$\begin{array}} \def\er{\end{array}{c} \\ \alpha} \def\be{\beta} \def\ga{\gamma + \be = 0,\;\; \ga+\be-\alpha} \def\be{\beta} \def\ga{\gamma+1=0,\;\; v(b)=1,\;\;w(b)=1/\omega \\ \\ \er $\hfill (C\,1) $\hspace*{-6mm}$are fulfilled. First, using relations (\ref{4.5}), (\ref{4.8}), (\ref{5.10}) and conditions (C\,1), we check \[\begin{array}} \def\er{\end{array}{l} \wh{U}_1 \wh{U}_2 = a\,(b+a^{-1}c\,d_0)\, e^{-i\xi}\;c\,e^{i\xi}= q^{-1/2+\be/2} c\,a(b+ q\,a^{-1}c\,d_0) \,e^{i\xi} \, e^{-i\xi} =\\ [1.5mm] = q^{\alpha} \def\be{\beta} \def\ga{\gamma/2+\be/2} c\,e^{i\xi}\,a(b+ q^{(\ga+\be-\alpha} \def\be{\beta} \def\ga{\gamma+1)/2}\,a^{-1}c\,d_0) \, e^{-i\xi} = \wh{U}_2 \wh{U}_1 . \er \] \[\begin{array}} \def\er{\end{array}{l} \wh{U}_1 \wh{U}_3 = a\,(b+a^{-1}c\,d_0)\,e^{-i\xi}\;c^{-1}\, (b^{-1}+q^{-1}\,a^{-1}c\,d_0)\, e^{-i\xi} = \\ [1.5mm] = q^{1/2-\be/2}c^{-1}\,a\, (b+q^{-1}\,a^{-1}c\,d_0)\,e^{-i\xi}\,(b^{-1}+q^{-1}\,a^{-1}c\,d_0)\, e^{-i\xi} =\\ [1.5mm] =q^{-\be/2}c^{-1}\,a\,(b^{-1}+q^{-1}\,a^{-1}c\,d_0)\, (b+q^{-1}\,a^{-1}c\,d_0)\,e^{-i\xi}\,e^{-i\xi} = \\ [1.5mm] = q^{-1/2-\be/2}c^{-1} \,(b^{-1}+q^{-1}\,a^{-1}c\,d_0)\,a\, (b+q^{-1}\,a^{-1}c\,d_0)\,e^{-i\xi}\, e^{-i\xi} = \\ [1.5mm] = q^{-1-\alpha} \def\be{\beta} \def\ga{\gamma/2-\be/2}c^{-1}\,(b^{-1}+q^{-1}\,a^{-1}c\,d_0)\,a\, e^{-i\xi}\,(b+a^{-1}c\,d_0)\,e^{-i\xi} = q^{-1} \wh{U}_3 \wh{U}_1 .\er \] The rest of relations (\ref{2.8}) can be proved similarly. Next, note that relation (\ref{2.10}) can be rewritten as follows \hspace*{3cm}$\begin{array}} \def\er{\end{array}{c} \\ \wh{U}_1\;\wh{U}_4\,(b-b^{-1})-\wh{U}_4\;\wh{U}_1\, (q\,b-q^{-1} \,b^{-1}) =-\omega\, \wh{U}_3 \; \wh{U}_2 \, b. \\ \\ \er $ \hfill (C\,2) $\hspace*{-7.5mm}$ To prove this equality we transform its l.h.s. and r.h.s. as follows \[\begin{array}} \def\er{\end{array}{l}\wh{U}_1\;\wh{U}_4\,(b-b^{-1})-\wh{U}_4\;\wh{U}_1\,(q\,b-q^{-1} \,b^{-1}) = a\,(b+a^{-1}c\,d_0)\,e^{-i\xi} \; a^{-1} e^{i\xi}\,(b-b^{-1})- \\ [1.5mm] - a^{-1} e^{i\xi} \; a\,(b+a^{-1}c\,d_0)\,e^{-i\xi} (q\,b-q^{-1}\,b^{-1}) = q^{-\alpha} \def\be{\beta} \def\ga{\gamma/2}\,(q^{1/2}\,b+q^{-1/2}\,a^{-1}c\,d_0)\,(b-b^{-1}) - \\ [1.5mm] - q^{-\alpha} \def\be{\beta} \def\ga{\gamma/2}\,(q^{-1/2}\,b+q^{1/2}\,a^{-1}c\,d_0)\,(q\,b-q^{-1}\,b^{-1}) = -q^{-\alpha} \def\be{\beta} \def\ga{\gamma/2}\,\omega\,(q^{1/2}\,a^{-1}c\,d_0\,b+q^{-1/2}); \\ \\ \wh{U}_3\;\wh{U}_2 \,b =c^{-1}\,(b^{-1}+q^{-1}\,a^{-1}c\,d_0)\, e^{-i\xi}\,c\,e^{i\xi} \,b = \\ [1.5mm] = q^{\be/2}\,(q^{-1/2}\,b^{-1}+q^{1/2}\, a^{-1}c\,d_0)\,b = q^{\be/2}\,(q^{-1/2}+q^{1/2}\,a^{-1}c\,d_0\,b). \er \] Thus, the equality (C\,2) is fulfilled if conditions (C\,1) are valid. The relation (\ref{2.11}) can be proved in the same way. \subsection*{\hspace*{6mm} \normalsize} \def\ds{\displaystyle\bf References.} \small \REF{1}{I.N.Bernstein, I.M.Gelfand, S.I.Gelfand, \ {\it Models for representations of Lie groups}. In: Proceedings of I.G.Petrovsky seminars, {\bf 2}, 3 (1976) (in Russian);\ I.N.Bernstein, I.M.Gelfand, S.I.Gelfand, \ Funct. analysis and its applications, {\bf 9}, No 4, 61 (1975). } \\ [1mm] \REF{2} {A.M.Polyakov, \ Mod.Phys.Lett. {\bf A 2}, 893 (1987).} \\[1mm] \REF{3} {A.Yu.Alekseev, S.L.Shatashvili, \ Comm. Math. Phys. {\bf 128}, 197 (1990); \\ \ H.La, P.Nelson, A.S.Schwarz, \ Comm. Math. Phys. {\bf 134}, 539 (1990).} \\ [1mm] \REF{4} {L.D.Faddeev, \ Comm. Math. Phys. {\bf 132}, 131 (1990). } \\ [1mm] \REF{5} {A.Yu.Alekseev, L.D.Faddeev, \ Comm. Math. Phys. {\bf 141}, 413 (1991).} \\ [1mm] \REF{6} {F.Falceto, K.Gawedzki, \ J. Geom. Phys. {\bf 11}, 251 (1993).} \\ [1mm] \REF{7} {A.A.Kirillov, \ {\it Elements of the theory of representations}. Springer-Verlag (1979);\\ \ A.Yu.Alekseev, L.D.Faddeev, S.L.Shatashvili, \ J. Geom. Phys. {\bf 5}, 391 (1989). } \\ [1mm] \REF{8} {A.Connes,\ {\it Noncommutative geometry}. (Academic Press, 1994).} \\ [1mm] \REF{9} {L.D.Faddeev, N.Yu.Reshetikhin, L.A.Takhtajan, \ Algebra and analysis {\bf 1}, 193 (1990). } \\ [1mm] \REF{10} {A.Yu.Alekseev, L.D.Faddeev, M.A.Semenov-Tian-Shansky, A.Yu.Volkov, \ {\it The unraveling of the quantum group structure in the WZNW theory}, preprint CERN-TH-5981/91 (1991). } \\ [1mm] \REF{11} {N.Yu.Reshetikhin, M.A.Semenov-Tian-Shansky, \ Lett. Math. Phys. {\bf 19}, 133 (1990). } \\ [1mm] \REF{12} {P.P.Kulish, N.Yu.Reshetikhin, \ Zapiski Nauch.Semin. LOMI {\bf 101}, 101 (1981) (in Russian). } \\ [1mm] \REF{13} {A.G.Bytsko, V.Schomerus,\ {\it Vertex operators -- from toy model to lattice algebras}, preprint q-alg/9611010 (1996).} \\ [1mm] \REF{14} {A.P. Isaev,\ J. Phys. {\bf A 29}, 6903 (1996).} \\ [1mm] \REF{15} {J-L.Gervais, A.Neveu, \ Nucl. Phys. {\bf B 238}, 125 (1984);\\ \ E.Cremmer, J-L.Gervais, \ Comm. Math. Phys. {\bf 134}, 619 (1990). } \\ [1mm] \REF{16} {O.Babelon, \ Comm. Math. Phys. {\bf 139}, 619 (1991).} \\ [1mm] \REF{17} {J.Avan, O.Babelon, E.Billey,\ Comm. Math. Phys. {\bf 178}, 281 (1996). } \\ [1mm] \REF{18} {A.N.Kirillov, N.Yu.Reshetikhin, \ Adv. Series in Math. Phys. {\bf 11}, 202 (World Scientific, 1990).} \\ [1mm] \REF{19} {A.O.Barut, R.Raczka, \ {\it Theory of group representations and applications} (Scient. Publishers, 1977);\\ \ L.C.Biedenharn, J.D.Louck, \ {\it Angular momentum in quantum physics}, Encyclopedia of mathematics and its applications, {\bf v.8} (Addison-Wesley, 1981). \\ [1mm] \REF{20} {M.Arik, D.D.Coon, \ J. Math. Phys. {\bf 17}, 524 (1976);\\ \ L.C.Biedenharn, \ J. Phys. {\bf A 22}, L873 (1989);\\ \ A.J.Macfarlane, \ J. Phys. {\bf A 22}, 4581 (1989);\\ \ P.P.Kulish, E.Damaskinsky, \ J. Phys. {\bf A 23}, L415 (1990);\\ \ M.Chaichian, P.P.Kulish, \ Phys. Lett. {\bf B 234}, 72 (1990).} \\ [1mm] \REF{21} {M.Nomura. \ J. Math. Phys. {\bf 30}, 2397 (1990);\\ \ L.Vaksman. \ Sov. Math. Dokl. {\bf 39}, 467 (1989);\\ \ H. Ruegg. \ J. Math. Phys. {\bf 31}, 1085 (1990). } \\ [1mm] \REF{22} {L.C.Biedenharn, M.Tarlini, \ Lett. Math. Phys. {\bf 20}, 271 (1990);\\ \ K.Bragiel, \, Lett. Math. Phys. {\bf 21}, 181 (1991);\\ \ G.Mack, V.Schomerus,\, Phys. Lett. {\bf B 267}, 207 (1991);\\ \ V.Rittenberg, M.Scheunert, \, J. Math. Phys. {\bf 33}, 436 (1992).} \\ [1mm] \REF{23} {A.G.Bytsko,\ {\it Tensor operators in $R$-matrix approach}, preprint DESY-95-254 (1995).} \end{document}
\section*{Preface} The {\tsc{Pythia}} and {\tsc{Jetset}} programs are frequently used for event generation in high-energy physics. The emphasis is on multiparticle production in collisions between elementary particles. This in particular means hard interactions in $\e^+\e^-$, $\p\p$ and $\e\p$ colliders, although also other applications are envisaged. The programs are intended to generate complete events, in as much detail as experimentally observable ones, within the bounds of our current understanding of the underlying physics. Many of the components of the programs represent original research, in the sense that models have been developed and implemented for a number of aspects not covered by standard theory. Although originally conceived separately, the {\tsc{Pythia}} and {\tsc{Jetset}} programs today are so often used together that it makes sense to present them here without too much distinction. Both programs have a long history, and several manuals have come out. The former round of {\tsc{Pythia/Jetset}} program descriptions appeared in 1987. Meanwhile a large number of additions and changes have been made. Recently a new description therefore appeared in \\[1mm] T. Sj\"ostrand, Computer Physics Commun. {\bf 82} (1994) 74. \\[1mm] This is the one and only correct reference to the current versions of {\tsc{Pythia}} and {\tsc{Jetset}}. The long writeup that you now have before you is an (unpublished) appendix to the publication above, and need not be separately cited. Instead remember to cite the original literature on the physics topics of particular relevance for your studies. (There is no reason to omit references to good physics papers simply because some of their contents have also been made available as program code.) Event generators often have a reputation for being `black boxes'; if nothing else, this report should provide you with a glimpse of what goes on inside the programs. Some such understanding may be of special interest for new users, who have no background in the field. An attempt has been made to structure the report sufficiently well that many of the sections can be read independently of each other, so you can pick the sections that interest you. I have tried to keep together the physics and the manual sections on specific topics, where practicable, which represents a change of policy compared with previous manual versions. Any feedback on this and other aspects is welcome. A large number of persons should be thanked for their contributions. Hans-Uno Bengtsson is the originator of the {\tsc{Pythia}} program, and for many years we worked in parallel on its further development. Mats Bengtsson is the main author of the final-state parton-shower algorithm. Bo Andersson and G\"osta Gustafson are the originators of the Lund model, and strongly influenced the early development of the programs. Further comments on the programs have been obtained from users too numerous to be mentioned here, but who are all gratefully acknowledged. To write programs of this size and complexity would be impossible without a strong user feedback. The moral responsibility for any remaining errors clearly rests with me. However, kindly note that this is a `University World' product, distributed `as is', free of charge, without any binding guarantees. And always remember that the programs do not represent a dead collection of established truths, but rather one of many possible approaches to the problem of multiparticle production in high-energy physics, at the frontline of current research. Be critical! \newpage \mbox{} \cleardoublepage \tableofcontents \cleardoublepage \pagestyle{plain} \setcounter{page}{1} \section{Introduction} Multiparticle production is the most characteristic feature of current high-energy physics. Today, observed particle multiplicities are typically between ten and a hundred, and with future machines this range will be extended upwards. The bulk of the multiplicity is found in jets, i.e. in bunches of hadrons (or decay products of hadrons) produced by the hadronization of quarks and gluons. \subsubsection*{The Complexity of High-Energy Processes} To first approximation, all processes have a simple structure at the level of interactions between the fundamental objects of nature, i.e. quarks, leptons and gauge bosons. For instance, a lot can be understood about the structure of hadronic events at LEP just from the `skeleton' process $\e^+\e^- \to \mathrm{Z}^0 \to \mathrm{q} \overline{\mathrm{q}}$. Corrections to this picture can be subdivided, arbitrarily but conveniently, into three main classes. Firstly, there are bremsstrahlung-type modifications, i.e. the emission of additional final-state particles by branchings such as $\mathrm{e} \to \mathrm{e} \gamma$ or $\mathrm{q} \to \mathrm{q} \mathrm{g}$. Because of the largeness of the strong coupling constant $\alpha_{\mathrm{s}}$, and because of the presence of the triple gluon vertex, QCD emission off quarks and gluons is especially prolific. We therefore speak about `parton showers', wherein a single initial parton may give rise to a whole bunch of partons in the final state. Also photon emission may give sizeable effects in $\e^+\e^-$ and $\e\p$ processes. The bulk of the bremsstrahlung corrections are universal, i.e. do not depend on the details of the process studied, but only on one or a few key numbers, such as the momentum transfer scale of the process. Such universal corrections may be included to arbitrarily high orders, using a probabilistic language. Alternatively, exact calculations of bremsstrahlung corrections may be carried out order by order in perturbation theory, but rapidly the calculations then become prohibitively complicated and the answers correspondingly lengthy. Secondly, we have `true' higher-order corrections, which involve a combination of loop graphs and the soft parts of the bremsstrahlung graphs above, a combination needed to cancel some divergences. In a complete description it is therefore not possible to consider bremsstrahlung separately, as assumed here. The necessary perturbative calculations are usually very difficult; only rarely have results been presented that include more than one non-`trivial' order, i.e. more than one loop. As above, answers are usually very lengthy, but some results are sufficiently simple to be generally known and used, such as the running of $\alpha_{\mathrm{s}}$, or the correction factor $1 + \alpha_{\mathrm{s}}/\pi + \cdots$ in the partial widths of $\mathrm{Z}^0 \to \mathrm{q} \overline{\mathrm{q}}$ decay channels. For high-precision studies it is imperative to take into account the results of loop calculations, but usually effects are minor for the qualitative aspects of high-energy processes. Thirdly, quarks and gluons are confined. In the two points above, we have used a perturbative language to describe the short-distance interactions of quarks, leptons and gauge bosons. For leptons and colourless bosons this language is sufficient. However, for quarks and gluons it must be complemented with a picture for the hadronization process (which can be subdivided into fragmentation and decays), wherein the coloured partons are transformed into jets of colourless hadrons, photons and leptons. This process is still not yet understood from first principles, but has to be based on models. In one sense, hadronization effects are overwhelmingly large, since this is where the bulk of the multiplicity comes from. In another sense, the overall energy flow of a high-energy event is mainly determined by the perturbative processes, with only a minor additional smearing caused by the hadronization step. One may therefore pick different levels of ambition, but in general detailed studies require a detailed modelling of the hadronization process. The simple structure that we started out with has now become considerably more complex --- instead of maybe two final-state partons we have a hundred final particles. The original physics is not gone, but the skeleton process has been dressed up and is no longer directly visible. A direct comparison between theory and experiment is therefore complicated at best, and impossible at worst. \subsubsection*{Event Generators} It is here that event generators come to the rescue. In an event generator, the objective strived for is to use computers to generate events as detailed as could be observed by a perfect detector. This is not done in one step, but rather by `factorizing' the full problem into a number of components, each of which can be handled reasonably accurately. Basically, this means that the hard process is used as input to generate bremsstrahlung corrections, and that the result of this exercise is thereafter left to hadronize. This sounds a bit easier than it really is --- else this report would be a lot thinner. However, the basic idea is there: if the full problem is too complicated to be solved in one go, try to subdivide it into smaller tasks of manageable proportions. In the actual generation procedure, most steps therefore involve the branching of one object into two, or at least into a very small number, each of which being free to branch in its turn. A lot of bookkeeping is involved, but much is of a repetitive nature, and can therefore be left for the computer to handle. As the name indicates, the output of an event generator should be in the form of `events', with the same average behaviour and the same fluctuations as real data. In the data, fluctuations arise from the quantum mechanics of the underlying theory. In generators, Monte Carlo techniques are used to select all relevant variables according to the desired probability distributions, and thereby ensure randomness in the final events. Clearly some loss of information is entailed: quantum mechanics is based on amplitudes, not probabilities. However, only very rarely do (known) interference phenomena appear that cannot be cast in a probabilistic language. This is therefore not a more restraining approximation than many others. Once there, an event generator can be used in many different ways. The five main applications are probably the following: \begin{Itemize} \item To give physicists a feeling for the kind of events one may expect/hope to find, and at what rates. \item As a help in the planning of a new detector, so that detector performance is optimized, within other constraints, for the study of interesting physics scenarios. \item As a tool for devising the analysis strategies that should be used on real data, so that signal-to-background conditions are optimized. \item As a method for estimating detector acceptance corrections that have to be applied to raw data, in order to extract the `true' physics signal. \item As a convenient framework within which to interpret the observed phenomena in terms of a more fundamental underlying theory (usually the Standard Model). \end{Itemize} Where does a generator fit into the overall analysis chain of an experiment? In `real life', the machine produces interactions. These events are observed by detectors, and the interesting ones are written to tape by the data acquisition system. Afterwards the events may be reconstructed, i.e. the electronics signals (from wire chambers, calorimeters, and all the rest) may be translated into a deduced setup of charged tracks or neutral energy depositions, in the best of worlds with full knowledge of momenta and particle species. Based on this cleaned-up information, one may proceed with the physics analysis. In the Monte Carlo world, the r\^ole of the machine, namely to produce events, is taken by the event generators described in this report. The behaviour of the detectors --- how particles produced by the event generator traverse the detector, spiral in magnetic fields, shower in calorimeters, or sneak out through cracks, etc. --- is simulated in programs such as \tsc{Geant} \cite{Bru89}. Traditionally, this latter activity is called event simulation, which is somewhat unfortunate since the same words could equally well be applied to what, here, we call event generation. A more appropriate term is detector simulation. Ideally, the output of this simulation has exactly the same format as the real data recorded by the detector, and can therefore be put through the same event reconstruction and physics analysis chain, except that here we know what the `right answer' should be, and so can see how well we are doing. Since the full chain of detector simulation and event reconstruction is very time-consuming, one often does `quick and dirty' studies in which these steps are skipped entirely, or at least replaced by very simplified procedures which only take into account the geometric acceptance of the detector and other trivial effects. One may then use the output of the event generator directly in the physics studies. There are still many holes in our understanding of the full event structure, despite an impressive amount of work and detailed calculations. To put together a generator therefore involved making a choice on what to include, and how to include it. At best, the spread between generators can be used to give some impression of the uncertainties involved. A multitude of approximations will be discussed in the main part of this report, but already here is should be noted that many major approximations are related to the almost complete neglect of the second point above, i.e. of the non-`trivial' higher-order effects. It can therefore only be hoped that the `trivial' higher order parts give the bulk of the experimental behaviour. By and large, this seems to be the case; for $\e^+\e^-$ annihilation it even turns out to be a very good approximation. The necessity to make compromises has one major implication: to write a good event generator is an art, not an exact science. It is therefore essential not to blindly trust the results of any single event generator, but always to make several cross-checks. In addition, with computer programs of tens of thousands of lines, the question is not whether bugs exist, but how many there are, and how critical their positions. Further, an event generator cannot be thought of as all-powerful, or able to give intelligent answers to ill-posed questions; sound judgement and some understanding of a generator are necessary prerequisites for successful use. In spite of these limitations, the event generator approach is the most powerful tool at our disposal if we wish to gain a detailed and realistic understanding of physics at current or future high-energy colliders. \subsubsection*{The Origins of the JETSET and PYTHIA Programs} Over the years, many event generators have appeared. Surveys of generators for $\e^+\e^-$ physics in general and LEP in particular may be found in \cite{Kle89,Sjo89}, for high-energy hadron--hadron ($\p\p$) physics in \cite{Ans90,Sjo92,Kno93}, and for $\e\p$ physics in \cite{HER92}. We refer the reader to those for additional details and references. In this particular report, the two closely connected programs {\tsc{Jetset}} and {\tsc{Pythia}} will be described. {\tsc{Jetset}} has its roots in the efforts of the Lund group to understand the hadronization process, starting in the late seventies \cite{And83}. The so-called string fragmentation model was developed as an explicit and detailed framework, within which the long-range confinement forces are allowed to distribute the energies and flavours of a parton configuration among a collection of primary hadrons, which subsequently may decay further. This model, known as the Lund string model, or `Lund' for short, contained a number of specific predictions, which were confirmed by data from PETRA and PEP, whence the model gained a widespread acceptance. The Lund string model is still today the most elaborate and widely used fragmentation model at our disposal. It remains at the heart of the {\tsc{Jetset/Pythia}} programs. In order to predict the shape of events at PETRA/PEP, and to study the fragmentation process in detail, it was necessary to start out from the partonic configurations that were to fragment. The generation of complete $\e^+\e^-$ hadronic events was therefore added, originally based on simple $\gamma$ exchange and first-order QCD matrix elements, later extended to full $\gamma^* / \Z^0$ exchange with first-order initial-state QED radiation and second-order QCD matrix elements. A number of utility routines were also provided early on, for everything from event listing to jet finding. By the mid-eighties it was clear that the matrix-element approach had reached the limit of its usefulness, in the sense that it could not fully describe the multijet topologies of the data. (Later on, the use of optimized perturbation theory was to lead to a resurgence of the matrix-element approach, but only for specific applications.) Therefore a parton-shower description was developed \cite{Ben87a} as an alternative to the matrix-element one. The combination of parton showers and string fragmentation has been very successful, and forms the main approach to the description of hadronic $\mathrm{Z}^0$ events. In recent years, {\tsc{Jetset}} has been a fairly stable product, covering the four main areas of fragmentation, final-state parton showers, $\e^+\e^-$ event generation and general utilities. The successes of string fragmentation in $\e^+\e^-$ made it interesting to try to extend this framework to other processes, and explore possible physics consequences. Therefore a number of other programs were written, which combined a process-specific description of the hard interactions with the general fragmentation framework of {\tsc{Jetset}}. The {\tsc{Pythia}} program evolved out of early studies on fixed-target proton--proton processes, addressed mainly at issues related to string drawing. With time, the interest shifted towards higher energies, first to the SPS $\p\pbar$ collider, and later to SSC and LHC, in the context of a number of workshops in the USA and Europe. Parton showers were added, for final-state radiation by making use of the {\tsc{Jetset}} routine, for initial-state one by the development of the concept of `backwards evolution', specifically for {\tsc{Pythia}} \cite{Sjo85}. Also a framework was developed for minimum-bias and underlying events \cite{Sjo87a}. Another main change was the introduction of an increasing number of hard processes, within the Standard Model and beyond. A special emphasis was put on the search for the Standard Model Higgs, in different mass ranges and in different channels, with due respect to possible background processes. The bulk of the machinery developed for hard processes actually depended little on the choice of initial state, as long as the appropriate parton distributions were there for the incoming partons and particles. It therefore made sense to extend the program from being only a $\p\p$ generator to working also for $\e^+\e^-$ and $\e\p$. This process was only completed in 1991, again spurred on by physics workshop activities. Currently {\tsc{Pythia}} should therefore work equally well for a selection of different possible incoming beam particles. The tasks of including new processes, and of improving the simulation of already present ones, are never-ending. Work therefore continues apace. While {\tsc{Jetset}} still is formally independent of {\tsc{Pythia}}, their ties have grown much stronger over the years, and the border-line between the two programs has become more and more artificial. It is no coincidence that the two are presented together here; this way a lot of repetition of common material can be avoided. The price to be paid is that some differences in philosophy will have to be discussed. \subsubsection*{About this Report} As we see, {\tsc{Jetset}} and {\tsc{Pythia}} started out as very ideologically motivated programs, developed to study specific physics questions in enough detail that explicit predictions could be made for experimental quantities. As it was recognized that experimental imperfections could distort the basic predictions, the programs were made available for general use by experimentalists. It thus became feasible to explore the models in more detail than would otherwise have been possible. As time went by, the emphasis came to shift somewhat, away from the original strong coupling to a specific fragmentation model, towards a description of high-energy multiparticle production processes in general. Correspondingly, the use expanded from being one of just comparing data with specific model predictions, to one of extensive use for the understanding of detector performance, for the derivation of acceptance correction factors, for the prediction of physics at future high-energy accelerators, and for the design of related detectors. While the ideology may be less apparent, it is still there, however. This is not something unique to the programs discussed here, but inherent in any event generator, or at least any generator that attempts to go beyond the simple parton level skeleton description of a hard process. Do not accept the myth that everything available in Monte Carlo form represents ages-old common knowledge, tested and true. Ideology is present by commissions or omissions in any number of details. Programs like {\tsc{Pythia}} and {\tsc{Jetset}} represent a major amount of original physics research, often on complicated topics where no simple answers are available. As a (potential) program user you must be aware of this, so that you can form your own opinion, not just about what to trust and what not to trust, but also how much to trust a given prediction, i.e. how uncertain it is likely to be. {\tsc{Jetset}} and {\tsc{Pythia}} are particularly well endowed in this respect, since a number of publications exist where most of the relevant physics is explained in considerable detail. In fact, the problem may rather be the opposite, to find the relevant information among all the possible places. One main objective of the current report is therefore to collect much of this information in one single place. Not all the material found in specialized papers is reproduced, by a wide margin, but at least enough should be found here to understand the general picture and to know where to go for details. The current report is therefore intended to replace the previous round of published physics descriptions and program manuals \cite{Sjo86,Sjo87,Ben87}. The formal new standard reference is \cite{Sjo94}, which is a fairly brief summary of this report --- for obvious reasons the full description is too long to be published in its entirety. Further specification could include a statement of the type `We use {\tsc{Pythia}} version X.x and {\tsc{Jetset}} version Y.y'. (If you are a {\LaTeX} fan, you may want to know that the program names in this report have been generated by the commands \verb+\textsc{Jetset}+ and \verb+\textsc{Pythia}+.) Kindly do not refer to {\tsc{Jetset/Pythia}} as `unpublished', `private communication' or `in preparation': such phrases are only creating unnecessary confusion. In addition, remember that many of the individual physics components are documented in separate publications. If some of these contain ideas that are useful to you, there is every reason to cite them. A reasonable selection would vary as a function of the physics you are studying. The criterion for which to pick should be simple: imagine that a Monte Carlo implementation had not been available. Would you then have cited a given paper on the grounds of its physics contents alone? If so, do not punish the extra effort of turning these ideas into publicly available software. (Monte Carlo manuals are good for nothing in the eyes of many theorists, so often only the acceptance of `mainstream' publications counts.) Here follows a list of some main areas where the $\tsc{Pythia/Jetset}$ programs contain original research: \begin{Itemize} \item The string fragmentation model \cite{And83}. \item The string effect \cite{And80}. \item Baryon production (diquark/popcorn) \cite{And82,And85}. \item Fragmentation of multiparton systems \cite{Sjo84}. \item Fragmentation effects on $\alpha_{\mathrm{s}}$ determinations \cite{Sjo84a}. \item Initial state parton showers \cite{Sjo85}. \item Final state parton showers \cite{Ben87a}. \item Photon radiation from quarks \cite{Sjo92c} \item Deep inelastic scattering \cite{And81a,Ben88}. \item Photoproduction \cite{Sch93a} and $\gamma\gamma$ physics \cite{Sch94a}. \item Parton distributions of the photon \cite{Sch95}. \item Colour flow in hard scatterings \cite{Ben84}. \item Elastic and diffractive cross sections \cite{Sch94}. \item Minijets (multiple parton--parton interactions) \cite{Sjo87a}. \item Rapidity gaps \cite{Dok92}. \item Jet clustering in $k_{\perp}$ \cite{Sjo83}. \end{Itemize} In addition to a physics survey, the current report also contains a complete manual for the two programs. Such manuals have always been updated and distributed jointly with the programs. To a first approximation, we therefore do not have much new to offer here. However, an attempt has been made to group the material more logically according to physics topics than in previous distributions, to tie it closer to the physics description, and to improve the layout and therefore the readability. Any feedback is welcome. A word of warning may be in place. The program description is fairly lengthy, and certainly could not be absorbed in one sitting. This is not even necessary, since all switches and parameters are provided with sensible default values, based on our best understanding (of the physics, and of what you expect to happen if you do not specify any options). As a new user, you can therefore disregard all the fancy options, and just run the program with a minimum ado. Later on, as you gain experience, the options that seem useful can be tried out. No single user is ever likely to find need for more than a fraction of the total number of possibilities available, yet many of them have been added to meet specific user requests. In some instances, not even this report will provide you with all the information you desire. You may wish to find out about recent versions of the program, know about related software, pick up a few sample main programs to get going, or get hold of related physics papers. Some such material can be found if you link to my World Wide Web homepage:\\ \ttt{http://thep.lu.se/tf2/staff/torbjorn/Welcome.html}\\ and study the contents there. \subsubsection*{Disclaimer} At all times it should be remembered that this is not a commercial product, developed and supported by professionals. Instead it is a `University World' product, developed by a very few physicists (mainly the current author) originally for their own needs, and supplied to other physicists on an `as-is' basis, free of charge. No guarantees are therefore given for the proper functioning of the programs, nor for the validity of physics results. In the end, it is always up to you to decide for yourself whether to trust a given result or not. Usually this requires comparison either with analytical results or with results of other programs, or with both. Even this is not necessarily foolproof: for instance, if an error is made in the calculation of a matrix element for a given process, this error will be propagated both into the analytical results based on the original calculation and into all the event generators which subsequently make use of the published formulae. In the end, there is no substitute for a sound physics judgement. This does not mean that you are all on your own, with a program nobody feels responsible for. Attempts are made to check processes as carefully as possible, to write programs that do not invite unnecessary errors, and to provide a detailed and accurate documentation. All of this while maintaining the full power and flexibility, of course, since the physics must always take precedence in any conflict of interests. If nevertheless any errors or unclarities are found, please do communicate them to me, e.g. on phone +46 -- 46 -- 222 48 16 or e-mail <EMAIL>. Every attempt will be made to solve problems as soon as is reasonably possible, given that this support is by one person alone, who also has other responsibilities. \subsubsection*{Appendix: The Historical Pythia} While the origin and connotations of the `{\tsc{Jetset}}' program name should be commonly known, the `{\tsc{Pythia}}' label may need some explanation. The myth tells how Apollon, the God of Wisdom, killed the powerful dragon-like monster Python, close to the village of Delphi in Greece. To commemorate this victory, Apollon founded the Pythic Oracle in Delphi, on the slopes of Mount Parnassos. Here men could come to learn the will of the Gods and the course of the future. The oracle plays an important r\^ole in many of the other Greek myths, such as those of Heracles and of King Oedipus. Questions were to be put to the Pythia, the `Priestess' or `Prophetess' of the Oracle. In fact, she was a local woman, usually a young maiden, of no particular religious schooling. Seated on a tripod, she inhaled the obnoxious vapours that seeped up through a crevice in the ground. This brought her to a trance-like state, in which she would scream seemingly random words and sounds. It was the task of the professional priests in Delphi to record those utterings and edit them into the official Oracle prophecies, which often took the form of poems in perfect hexameter. In fact, even these edited replies were often less than easy to interpret. The Pythic oracle acquired a reputation for ambiguous answers. The Oracle existed already at the beginning of the historical era in Greece, and was universally recognized as the foremost religious seat. Individuals and city states came to consult, on everything from cures for childlessness to matters of war. Lavish gifts allowed the temple area to be built and decorated. Many states supplied their own treasury halls, where especially beautiful gifts were on display. Sideshows included the Omphalos, a stone reputedly marking the centre of the Earth, and the Pythic games, second only to the Olympic ones in importance. Strife inside Greece eventually led to a decline in the power of the Oracle. A serious blow was dealt when the Oracle of Zeus Ammon (see below) declared Alexander the Great to be the son of Zeus. The Pythic Oracle lived on, however, and was only closed by a Roman Imperial decree in 390 \tsc{ad}, at a time when Christianity was ruthlessly destroying any religious opposition. Pythia then had been at the service of man and Gods for a millenium and a half. The r\^ole of the Pythic Oracle replies on the course of history is nowhere better described than in `The Histories' by Herodotus \cite{HerBC}, the classical and captivating description of the Ancient World at the time of the Great War between Greeks and Persians. Especially famous is the episode with King Croisus of Lydia. Contemplating a war against the upstart Persian Empire, he resolves to ask an oracle what the outcome of a potential battle would be. However, to have some guarantee for the veracity of any prophecy, he decides to send embassies to all the renowned oracles of the known World. The messengers are instructed to inquire the various divinities, on the hundredth day after their departure, what King Croisus is doing at that very moment. From the Pythia the messengers bring back the reply \begin{em}\begin{verse} I know the number of grains of sand as well as the expanse of the sea, \\ And I comprehend the dumb and hear him who does not speak, \\ There came to my mind the smell of the hard-shelled turtle, \\ Boiled in copper together with the lamb, \\ With copper below and copper above. \end{verse}\end{em} The veracity of the Pythia is thus established by the crafty ruler, who had waited until the appointed day, slaughtered a turtle and a lamb, and boiled them together in a copper cauldron with a copper lid. Also the Oracle of Zeus Ammon in the Libyan desert is able to give a correct reply (lost to posterity), while all others fail. King Croisus now sends a second embassy to Delphi, inquiring after the outcome of a battle against the Persians. The Pythia answers \begin{em}\begin{verse} If Croisus passes over the Halys he will dissolve a great Empire. \end{verse}\end{em} Taking this to mean he would win, the King collects his army and crosses the border river, only to suffer a crushing defeat and see his Kingdom conquered. When the victorious King Cyrus allows Croisus to send an embassy to upbraid the Oracle, the God Apollon answers through his Prophetess that he has correctly predicted the destruction of a great empire --- Croisus' own --- and that he cannot be held responsible if people choose to interpret the Oracle answers to their own liking. The history of the {\tsc{Pythia}} program is neither as long nor as dignified as that of its eponym. However, some points of contact exist. You must be very careful when you formulate the questions: any ambiguities will corrupt the reply you get. And you must be even more careful not to misinterpret the answers; in particular not to pick the interpretation that suits you before considering the alternatives. Finally, even a perfect God has servants that are only human: a priest might mishear the screams of the Pythia and therefore produce an erroneous oracle reply; the current author might unwittingly let a bug free in the program {\tsc{Pythia}}. \clearpage \section{Physics Overview} In this section we will try to give an overview of the main physics features of {\tsc{Jetset}} and {\tsc{Pythia}}, and also to introduce some terminology. The details will be discussed in subsequent sections. For the description of a typical high-energy event, an event generator should contain a simulation of several physics aspects. If we try to follow the evolution of an event in some semblance of a time order, one may arrange these aspects as follows: \begin{Enumerate} \item Initially two beam particles are coming in towards each other. Normally each particle is characterized by a set of parton distribution functions, which defines the partonic substructure in terms of flavour composition and energy sharing. \item One shower initiator parton from each beam starts off a sequence of branchings, such as $\mathrm{q} \to \mathrm{q} \mathrm{g}$, which build up an initial-state shower. \item One incoming parton from each of the two showers enters the hard process, where then a number of outgoing partons are produced, usually two. It is the nature of this process that determines the main characteristics of the event. \item Also the outgoing partons may branch, to build up final-state showers. \item When a shower initiator is taken out of a beam particle, a beam remnant is left behind. This remnant may have an internal structure, and a net colour charge that relates it to the rest of the final state. \item The QCD confinement mechanism ensures that the outgoing quarks and gluons are not observable, but instead fragment to colour neutral hadrons. \item Many of the produced hadrons are unstable and decay further. \end{Enumerate} Conventionally, only quarks and gluons are counted as partons, while leptons and photons are not. If pushed {\it ad absurdum} this may lead to some unwieldy terminology. We will therefore, where it does not matter, speak of an electron or a photon in the `partonic' substructure of an electron, lump branchings $\mathrm{e} \to \mathrm{e} \gamma$ together with other `parton shower' branchings such as $\mathrm{q} \to \mathrm{q} \mathrm{g}$, and so on. With this notation, the division into the above seven points applies equally well to an interaction between two leptons, between a lepton and a hadron, and between two hadrons. In the following subsections, we will survey the above seven aspects, not in the same order as given here, but rather in the order in which they appear in the program execution, i.e. starting with the hard process. \subsection{Hard Processes and Parton Distributions} In {\tsc{Jetset}}, only two hard processes are available. The first and main one is $\e^+\e^- \to \gamma^* / \Z^0 \to \mathrm{q} \overline{\mathrm{q}}$. Here the `$*$' of $\gamma^*$ is used to denote that the photon must be off the mass shell. The distinction is of some importance, since a photon on the mass shell cannot decay. Of course also the $\mathrm{Z}^0$ can be off the mass shell, but here the distinction is less relevant (strictly speaking, a $\mathrm{Z}^0$ is always off the mass shell). In the following we may not always use `$*$' consistently, but the rule of thumb is to use a `$*$' only when a process is not kinematically possible for a particle of nominal mass. The quark $\mathrm{q}$ in the final state of $\e^+\e^- \to \gamma^* / \Z^0 \to \mathrm{q} \overline{\mathrm{q}}$ may be $\u$, $\d$, $\mathrm{s}$, $\c$, $\b$ or $\t$; the flavour in each event is picked at random, according to the relative couplings, evaluated at the hadronic c.m. energy. Also the angular distribution of the final $\mathrm{q} \overline{\mathrm{q}}$ pair is included. No parton-distribution functions are needed. The other {\tsc{Jetset}} process is a routine to generate $\mathrm{g} \mathrm{g} \mathrm{g}$ and $\gamma \mathrm{g} \mathrm{g}$ final states, as expected in onium 1$^{--}$ decays such as $\Upsilon$. Given the current limits on the top mass, toponium will decay weakly much too fast for these processes to be of any interest, so therefore no new applications are expected. {\tsc{Pythia}} contains a much richer selection, with close to a hundred different hard processes. These may be classified in many different ways. One is according to the number of final-state objects: we speak of `$2 \to 1$' processes, `$2 \to 2$' ones, `$2 \to 3$' ones, etc. This aspect is very relevant from a programming point of view: the more particles in the final state, the more complicated the phase space and therefore the whole generation procedure. In fact, {\tsc{Pythia}} is optimized for $2 \to 1$ and $2 \to 2$ processes. There is currently no generic treatment of processes with three or more particles in the final state, but rather a few different machineries, each tailored to the pole structure of a specific class of graphs. This may be seen as a major limitation, and indeed is so at times. However, often one can come quite far with only one or two particles in the final state, since showers will add the required extra activity. The classification may also be misleading at times, since an $s$-channel resonance is considered as a single particle, even if it is assumed always to decay into two final-state particles. Thus the process $\e^+\e^- \to \mathrm{W}^+ \mathrm{W}^- \to \mathrm{q}_1 \overline{\mathrm{q}}'_1 \, \mathrm{q}_2 \overline{\mathrm{q}}'_2$ is classified as $2 \to 2$, although the decay treatment of the $\mathrm{W}$ pair includes the full $2 \to 4$ matrix elements. Another classification is according to the physics scenario. This will be the main theme of section \ref{s:pytproc}. The following major groups may be distinguished: \begin{Itemize} \item Hard QCD processes, e.g. $\mathrm{q} \mathrm{g} \to \mathrm{q} \mathrm{g}$. \item Soft QCD processes, such as diffractive and elastic scattering, and minimum-bias events. \item Heavy-flavour production, e.g. $\mathrm{g} \mathrm{g} \to \t \overline{\mathrm{t}}$. \item Prompt-photon production, e.g. $\mathrm{q} \mathrm{g} \to \mathrm{q} \gamma$. \item Photon-induced processes, e.g. $\gamma \mathrm{g} \to \mathrm{q} \overline{\mathrm{q}}$. \item Deep inelastic scattering, e.g. $\mathrm{q} \ell \to \mathrm{q} \ell$. \item $\mathrm{W} / \mathrm{Z}$ production, such as the $\e^+\e^- \to \gamma^* / \Z^0$ already found in {\tsc{Jetset}}, or $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{W}^+ \mathrm{W}^-$. \item Standard model Higgs production, where the Higgs is reasonably light and narrow, and can therefore still be considered as a resonance. \item Gauge boson scattering processes, such as $\mathrm{W} \mathrm{W} \to \mathrm{W} \mathrm{W}$, when the Standard Model Higgs is so heavy and broad that resonant and non-resonant contributions have to be considered together. \item Non-standard Higgs particle production, within the framework of a two-Higgs-doublet scenario with three neutral and two charged Higgs states. \item Production of new gauge bosons, such as a $\mathrm{Z}'$. \item Production of fourth-generation fermions. \item Leptoquark production. \item Deviations from Standard Model processes, e.g. due to contact interactions or a strongly interacting gauge boson sector. These scenarios do not always appear as separate processes, but may just be options to some of the processes above. \end{Itemize} This is by no means a survey of all interesting physics. Most notable is the absence of supersymmetric particle production and decay, but many other examples could be found. Also, within the scenarios studied, not all contributing graphs have always been included, but only the more important and/or more interesting ones. In many cases, various approximations are involved in the matrix elements coded. The cross section for a given process $ij \to k$ is given by \begin{equation} \sigma_{ij \to k} = \int \d x_1 \int \d x_2 \, f^1_i(x_1) \, f^2_j(x_2) \, \hat{\sigma}_{ij \to k} ~. \end{equation} Here $\hat{\sigma}$ is the cross section for the hard partonic process, as codified in the matrix elements for each specific process. For processes with many particles in the final state it would be replaced by an integral over the allowed final-state phase space. The $f^a_i(x)$ are the parton-distribution functions, which describe the probability to find a parton $i$ inside beam particle $a$, with parton $i$ carrying a fraction $x$ of the total $a$ momentum. Actually, parton distributions also depend on some momentum scale $Q^2$ that characterizes the hard process. Parton distributions are most familiar for hadrons, such as the proton. Hadrons are inherently composite objects, made up of quarks and gluons. Since we do not understand QCD, a derivation from first principles of hadron parton distributions does not yet exist, although some progress is being made in lattice QCD studies. It is therefore necessary to rely on parametrizations, where experimental data are used in conjunction with the evolution equations for the $Q^2$ dependence, to pin down the parton distributions. Several different groups have therefore produced their own fits, based on slightly different sets of data, and with some variation in the theoretical assumptions. Also for fundamental particles, such as the electron, is it convenient to introduce parton distributions. The function $f^{\mathrm{e}}_{\mathrm{e}}(x)$ thus parametrizes the probability that the electron that takes part in the hard process retains a fraction $x$ of the original energy, the rest being radiated (into photons) in the initial state. Of course, such radiation could equally well be made part of the hard interaction, but the parton-distribution approach usually is much more convenient. If need be, a description with fundamental electrons is recovered for the choice $f_{\mathrm{e}}^{\mathrm{e}}(x, Q^2)=\delta(x-1)$. Note that, contrary to the proton case, electron parton distributions are calculable from first principles, and reduce to the $\delta$ function above for $Q^2 \to 0$. The electron may also contain photons, and the photon may in its turn contain quarks and gluons. The internal structure of the photon is a bit of a problem, since the photon contains a point-like part, which is perturbatively calculable, and a vector-meson dominance part, which is not. Normally, the photon parton distributions are therefore parametrized, just as the hadron ones. Since the electron ultimately contains quarks and gluons, hard QCD processes like $\mathrm{q} \mathrm{g} \to \mathrm{q} \mathrm{g}$ therefore not only appear in $\p\p$ collisions, but also in $\e\p$ ones (`resolved photoproduction') and in $\e^+\e^-$ ones (`doubly resolved 2$\gamma$ events'). The parton distribution function approach here makes it much easier to reuse one and the same hard process in different contexts. There is also another kind of possible generalization. The two processes $\mathrm{q} \overline{\mathrm{q}} \to \gamma^* / \Z^0$, studied in hadron colliders, and $\e^+\e^- \to \gamma^* / \Z^0$, studied in $\e^+\e^-$ colliders, are really special cases of a common process, $\mathrm{f} \overline{\mathrm{f}} \to \gamma^* / \Z^0$, where $\mathrm{f}$ denotes a fundamental fermion, i.e. a quark, lepton or neutrino. The whole structure is therefore only coded once, and then slightly different couplings and colour prefactors are used, depending on the initial state considered. Usually the interesting cross section is a sum over several different initial states, e.g. $\u \overline{\mathrm{u}} \to \gamma^* / \Z^0$ and $\d \overline{\mathrm{d}} \to \gamma^* / \Z^0$ in a hadron collider. This kind of summation is always implicitly done, even when not explicitly mentioned in the text. \subsection{Initial- and Final-State Radiation} In every process that contains coloured and/or charged objects in the initial or final state, gluon and/or photon radiation may give large corrections to the overall topology of events. Starting from a basic $2 \to 2$ process, this kind of corrections will generate $2 \to 3$, $2 \to 4$, and so on, final-state topologies. As the available energies are increased, hard emission of this kind is increasingly important, relative to fragmentation, in determining the event structure. Two traditional approaches exist to the modelling of perturbative corrections. One is the matrix-element method, in which Feynman diagrams are calculated, order by order. In principle, this is the correct approach, which takes into account exact kinematics, and the full interference and helicity structure. The only problem is that calculations become increasingly difficult in higher orders, in particular for the loop graphs. Only in exceptional cases have therefore more than one loop been calculated in full, and often we do not have any loop corrections at all at our disposal. On the other hand, we have indirect but strong evidence that, in fact, the emission of multiple soft gluons plays a significant r\^ole in building up the event structure, e.g. at LEP, and this sets a limit to the applicability of matrix elements. Since the phase space available for gluon emission increases with the available energy, the matrix-element approach becomes less relevant for the full structure of events at higher energies. However, the perturbative expansion by itself is better behaved at higher energies, owing to the running of $\alpha_{\mathrm{s}}$. As a consequence, inclusive measurements, e.g. of the rate of well-separated jets, should yield more reliable results. The second possible approach is the parton-shower one. Here an arbitrary number of branchings of one parton into two (or more) may be put together, to yield a description of multijet events, with no explicit upper limit on the number of partons involved. This is possible since the full matrix-element expressions are not used, but only approximations derived by simplifying the kinematics, and the interference and helicity structure. Parton showers are therefore expected to give a good description of the substructure of jets, but in principle the shower approach has limited predictive power for the rate of well-separated jets (i.e. the 2/3/4/5-jet composition). In practice, shower programs may be patched up to describe the hard-gluon emission region reasonably well, in particular for the $\e^+\e^-$ annihilation process. Nevertheless, the shower description is not optimal for absolute $\alpha_{\mathrm{s}}$ determinations. Thus the two approaches are complementary in many respects, and both have found use. However, because of its simplicity and flexibility, the parton-shower option is generally the first choice, while the matrix elements one is mainly used for $\alpha_{\mathrm{s}}$ determinations, angular distribution of jets, triple-gluon vertex studies, and other specialized studies. Obviously, the ultimate goal would be to have an approach where the best aspects of the two worlds are harmoniously married. \subsubsection{Matrix elements} Matrix elements are especially made use of in the {\tsc{Jetset}} implementation of the process $\e^+\e^- \to \gamma^* / \Z^0 \to \mathrm{q} \overline{\mathrm{q}}$. For initial-state QED radiation, a first order (unexponentiated) description has been adopted. This means that events are subdivided into two classes, those where a photon is radiated above some minimum energy, and those without such a photon. In the latter class, the soft and virtual corrections have been lumped together to give a total event rate that is correct up to one loop. This approach worked fine at PETRA/PEP energies, but does not do so well for the $\mathrm{Z}^0$ line shape, i.e. in regions where the cross section is rapidly varying and high precision is strived for. For final-state QCD radiation, several options are available. The default is the parton-shower one (see below), but the matrix-elements options are also frequently used. In the definition of 3- or 4-jet events, a cut is introduced whereby it is required that any two partons have an invariant mass bigger than some fraction of the c.m. energy. 3-jet events which do not fulfill this requirement are lumped with the 2-jet ones. The first-order matrix-element option, which only contains 3- and 2-jet events therefore involves no ambiguities. In second order, where also 4-jets have to be considered, a main issue is what to do with 4-jet events that fail the cuts. Depending on the choice of recombination scheme, whereby the two nearby partons are joined into one, different 3-jet events are produced. Therefore the second-order differential 3-jet rate has been the subject of some controversy, and {\tsc{Jetset}} actually contains two different implementations. By contrast, {\tsc{Pythia}} does not contain any full higher-order matrix elements, with loop contributions included. There are a few cases where higher-order matrix elements are included at the Born level. Consider e.g. the case of $\mathrm{W}$ production at a hadron collider, which is contained in the lowest-order process $\mathrm{q} \overline{\mathrm{q}}' \to \mathrm{W}$. In an inclusive description, additional jets recoiling against the $\mathrm{W}$ may be generated by parton showers. {\tsc{Pythia}} also contains the two first-order processes $\mathrm{q} \mathrm{g} \to \mathrm{W} \mathrm{q}'$ and $\mathrm{q} \overline{\mathrm{q}}' \to \mathrm{W} \mathrm{g}$. The cross sections for these processes are divergent when the $p_{\perp} \to 0$. In this region a correct treatment would therefore have to take into account loop corrections, which are not available in {\tsc{Pythia}}. Depending on the physics application, one could then use {\tsc{Pythia}} in one of two ways. In the region of small $p_{\perp}$, the preferred option is lowest-order matrix elements combined with parton showers. For the production of a $\mathrm{W}$ at large $p_{\perp}$, on the other hand, the shower approach is too imprecise to give the right cross section; additionally the event selection machinery is very inefficient. Here it is advantageous to generate first-order events, and then add showers only to describe additional softer radiation. \subsubsection{Parton showers} The separation of radiation into initial- and final-state showers is arbitrary, but very convenient. There are also situations where it is appropriate: for instance, the process $\e^+\e^- \to \mathrm{Z}^0 \to \mathrm{q} \overline{\mathrm{q}}$ only contains final-state QCD radiation (QED radiation, however, is possible both in the initial and final state), while $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{Z}^0 \to \e^+\e^-$ only contains initial-state QCD one. Similarly, the distinction of emission as coming either from the $\mathrm{q}$ or from the $\overline{\mathrm{q}}$ is arbitrary. In general, the assignment of radiation to a given mother parton is a good approximation for an emission close to the direction of motion of that parton, but not for the wide-angle emission in between two jets, where interference terms are expected to be important. In both initial- and final-state showers, the structure is given in terms of branchings $a \to bc$, specifically $\mathrm{e} \to \mathrm{e} \gamma$, $\mathrm{q} \to \mathrm{q} \mathrm{g}$, $\mathrm{q} \to \mathrm{q} \gamma$, $\mathrm{g} \to \mathrm{g} \mathrm{g}$, and $\mathrm{g} \to \mathrm{q} \overline{\mathrm{q}}$. Each of these processes is characterized by a splitting kernel $P_{a \to bc}(z)$. The branching rate is proportional to the integral $\int P_{a \to bc}(z) \, \d z$. The $z$ value picked for a branching describes the energy sharing, with daughter $b$ taking a fraction $z$ and daughter $c$ the remaining $1-z$ of the $a$ energy. Once formed, the daughters $b$ and $c$ may in turn branch, and so on. Each parton is characterized by some virtuality scale $Q^2$, which gives an approximate sense of time ordering to the cascade. In the initial-state shower, $Q^2$ values are gradually increasing as the hard scattering is approached, while $Q^2$ is decreasing in the final-state showers. Shower evolution is cut off at some lower scale $Q_0$, typically around 1 GeV for QCD branchings. The same cut-off scale is also used to regularize the soft gluon emission divergences in the splitting kernels. From above, a maximum scale $Q_{\mathrm{max}}$ is introduced, where the showers are matched to the hard interaction itself. The relation between $Q_{\mathrm{max}}$ and the kinematics of the hard scattering is uncertain, and the choice made can strongly affect the amount of well-separated jets. Despite a number of common traits, the initial- and final-state radiation machineries are in fact quite different, and are described separately below. For historical reasons, the final-state shower is found in {\tsc{Jetset}} and the initial-state one in {\tsc{Pythia}}. Final-state showers are time-like, i.e. partons have $m^2 = E^2 - \mbf{p}^2 \geq 0$. The evolution variable $Q^2$ of the cascade is therefore in {\tsc{Jetset}} associated with the $m^2$ of the branching parton, but this choice is not unique. Starting from $Q^2_{\mathrm{max}}$, an original parton is evolved downwards in $Q^2$ until a branching occurs. The selected $Q^2$ value defines the mass of the branching parton, and the $z$ of the splitting kernel the parton energy division between its daughters. These daughters may now, in turn, evolve downwards, in this case with maximum virtuality already defined by kinematics, and so on down to the $Q_0$ cut-off. In QCD showers, corrections to the leading-log picture, so-called coherence effects, lead to an ordering of subsequent emissions in terms of decreasing angles. This does not follow automatically from the mass-ordering constraint, but is implemented as an additional requirement on allowed emissions. Photon emission is not affected by angular ordering. It is also possible to obtain non-trivial correlations between azimuthal angles in the various branchings, some of which are implemented as options. Finally, the theoretical analysis strongly suggests the scale choice $\alpha_{\mathrm{s}} = \alpha_{\mathrm{s}}(p_{\perp}^2) = \alpha_{\mathrm{s}}(z(1-z)m^2)$, and this is the default in the program. The final-state radiation machinery is applied in the c.m. frame of the hard scattering. The total energy and momentum of the hard-scattering subsystem is preserved, as is the direction of the outgoing partons (in that frame). In contrast to final-state showers, initial-state ones are space-like. This means that, in the sequence of branchings $a \to bc$ that lead up from the shower initiator to the hard interaction, particles $a$ and $b$ have $m^2 = E^2 - \mbf{p}^2 <0$. The `side branch' particle $c$, which does not participate in the hard scattering, may be on the mass shell, or have a time-like virtuality. In the latter case a time-like shower will evolve off it, rather like the final-state radiation described above. To first approximation, the evolution of the space-like main branch is characterized by the evolution variable $Q^2 = -m^2$, which is required to be strictly increasing along the shower, i.e. $Q_b^2 > Q_a^2$. Corrections to this picture have been calculated, but are basically absent in {\tsc{Pythia}}. Initial-state radiation is handled within the backwards evolution scheme. In this approach, the choice of the hard scattering is based on the use of evolved parton distributions, which means that the inclusive effects of initial-state radiation are already included. What remains is therefore to construct the exclusive showers. This is done starting from the two incoming partons at the hard interaction, tracing the showers `backwards in time', back to the two shower initiators. In other words, given a parton $b$, one tries to find the parton $a$ that branched into $b$. The evolution in the Monte Carlo is therefore in terms of a sequence of decreasing space-like virtualities $Q^2$ and increasing momentum fractions $x$. Branchings on the two sides are interleaved in a common sequence of decreasing $Q^2$ values. In the above formalism, there is no real distinction between gluon and photon emission. Some of the details actually do differ, as will be explained in the full description. The initial- and final-state radiation shifts around the kinematics of the original hard interaction. In deep inelastic scattering, this means that the $x$ and $Q^2$ values that can be derived from the momentum of the scattered lepton do not agree with the values originally picked. In high-$p_{\perp}$ processes, it means that one no longer has two jets with opposite and compensating $p_{\perp}$, but more complicated topologies. Effects of any original kinematics selection cuts are therefore smeared out, an unfortunate side-effect of the parton-shower approach. \subsection{Beam Remnants} In a hadron--hadron collision, the initial-state radiation algorithm reconstructs one shower initiator in each beam. This initiator only takes some fraction of the total beam energy, leaving behind a beam remnant which takes the rest. For a proton beam, a $\u$ quark initiator would leave behind a $\u \d$ diquark beam remnant, with an antitriplet colour charge. The remnant is therefore colour-connected to the hard interaction, and forms part of the same fragmenting system. It is further customary to assign a primordial transverse momentum to the shower initiator, to take into account the motion of quarks inside the original hadron, basically as required by the uncertainty principle. This primordial $k_{\perp}$ is selected according to some suitable distribution, and the recoil is assumed to be taken up by the beam remnant. Often the remnant is more complicated, e.g. a $\mathrm{g}$ initiator would leave behind a $\u \u \d$ proton remnant system in a colour octet state, which can conveniently be subdivided into a colour triplet quark and a colour antitriplet diquark, each of which are colour-connected to the hard interaction. The energy sharing between these two remnant objects, and their relative transverse momentum, introduces additional degrees of freedom, which are not understood from first principles. Na\"{\i}vely, one would expect an $\e\p$ event to have only one beam remnant, and an $\e^+\e^-$ event none. This is not always correct, e.g. a $\gamma \gamma \to \mathrm{q} \overline{\mathrm{q}}$ interaction in an $\e^+\e^-$ event would leave behind the $\mathrm{e}^+$ and $\mathrm{e}^-$ as beam remnants, and a $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{g} \mathrm{g}$ interaction in resolved photoproduction in an $\e^+\e^-$ event would leave behind one $\mathrm{e}^{\pm}$ and one $\mathrm{q} / \overline{\mathrm{q}}$ in each remnant. Corresponding complications occur for photoproduction in $\e\p$ events. There is another source of beam remnants. If parton distributions are used to resolve an electron inside an electron, some of the original energy is not used in the hard interaction, but is rather associated with initial-state photon radiation. The initial-state shower is in principle intended to trace this evolution and reconstruct the original electron before any radiation at all took place. However, because of cut-off procedures, some small amount may be left unaccounted. Alternatively the user may have chosen to switch off initial-state radiation altogether, but still preserved the resolved electron parton distributions. In either case the remaining energy is given to a single photon of vanishing transverse momentum, which is then considered in the same spirit as `true' beam remnants. So far we have assumed that each event only contains one hard interaction, i.e. that each incoming particle has only one parton which takes part in hard processes, and that all other constituents sail through unaffected. This is appropriate in $\e^+\e^-$ or $\e\p$ events, but not necessarily so in hadron--hadron collisions. Here each of the beam particles contains a multitude of partons, and so the probability for several interactions in one and the same event need not be negligible. In principle these additional interactions could arise because one single parton from one beam scatters against several different partons from the other beam, or because several partons from each beam take place in separate $2 \to 2$ scatterings. Both are expected, but combinatorics should favour the latter, which is the mechanism considered in {\tsc{Pythia}}. The dominant $2 \to 2$ QCD cross sections are divergent for $p_{\perp} \to 0$, and drop rapidly for larger $p_{\perp}$. Probably the lowest-order perturbative cross sections will be regularized at small $p_{\perp}$ by colour coherence effects: an exchanged gluon of small $p_{\perp}$ has a large transverse wave function and can therefore not resolve the individual colour charges of the two incoming hadrons; it will only couple to an average colour charge that vanishes in the limit $p_{\perp} \to 0$. In the program, some effective $p_{\perp\mathrm{min}}$ scale is therefore introduced, below which the perturbative cross section is either assumed completely vanishing or at least strongly damped. Phenomenologically, $p_{\perp\mathrm{min}}$ comes out to be a number of the order of 1.5--2.0 GeV. In a typical `minimum-bias' event one therefore expects to find one or a few scatterings at scales around or a bit above $p_{\perp\mathrm{min}}$, while a high-$p_{\perp}$ event also may have additional scatterings at the $p_{\perp\mathrm{min}}$ scale. The probability to have several high-$p_{\perp}$ scatterings in the same event is small, since the cross section drops so rapidly with $p_{\perp}$. The understanding of multiple interaction is still very primitive, and even the experimental evidence that it exists at all is rather weak. {\tsc{Pythia}} therefore contains several different options, with a fairly simple one as default. The options differ in particular on the issue of the `pedestal' effect: is there an increased probability or not for additional interactions in an event which is known to contain a hard scattering, compared with one that contains no hard interactions? \subsection{Fragmentation} QCD perturbation theory, formulated in terms of quarks and gluons, is valid at short distances. At long distances, QCD becomes strongly interacting and perturbation theory breaks down. In this confinement regime, the coloured partons are transformed into colourless hadrons, a process called either hadronization or fragmentation. In this paper we reserve the former term for the combination of fragmentation and the subsequent decay of unstable particles. The fragmentation process has yet to be understood from first principles, starting from the QCD Lagrangian. This has left the way clear for the development of a number of different phenomenological models. Three main schools are usually distinguished, string fragmentation (SF), independent fragmentation (IF) and cluster fragmentation (CF), but many variants and hybrids exist. Being models, none of them can lay claims to being `correct', although some may be better founded than others. The best that can be aimed for is internal consistency, a good representation of existing data, and a predictive power for properties not yet studied or results at higher energies. {\tsc{Jetset}} is intimately connected with string fragmentation, in the form of the time-honoured `Lund model'. This is the default for all {\tsc{Jetset/Pythia}} applications, but independent fragmentation options also exist, for applications where one wishes to study the importance of string effects. All current models are of a probabilistic and iterative nature. This means that the fragmentation process as a whole is described in terms of one or a few simple underlying branchings, of the type jet $\to$ hadron + remainder-jet, string $\to$ hadron + remainder-string, and so on. At each branching, probabilistic rules are given for the production of new flavours, and for the sharing of energy and momentum between the products. To understand fragmentation models, it is useful to start with the simplest possible system, a colour-singlet $\mathrm{q} \overline{\mathrm{q}}$ 2-jet event, as produced in $\e^+\e^-$ annihilation. Here lattice QCD studies lend support to a linear confinement picture (in the absence of dynamical quarks), i.e. the energy stored in the colour dipole field between a charge and an anticharge increases linearly with the separation between the charges, if the short-distance Coulomb term is neglected. This is quite different from the behaviour in QED, and is related to the presence of a triple-gluon vertex in QCD. The details are not yet well understood, however. The assumption of linear confinement provides the starting point for the string model. As the $\mathrm{q}$ and $\overline{\mathrm{q}}$ partons move apart from their common production vertex, the physical picture is that of a colour flux tube (or maybe colour vortex line) being stretched between the $\mathrm{q}$ and the $\overline{\mathrm{q}}$. The transverse dimensions of the tube are of typical hadronic sizes, roughly 1 fm. If the tube is assumed to be uniform along its length, this automatically leads to a confinement picture with a linearly rising potential. In order to obtain a Lorentz covariant and causal description of the energy flow due to this linear confinement, the most straightforward way is to use the dynamics of the massless relativistic string with no transverse degrees of freedom. The mathematical, one-dimensional string can be thought of as parametrizing the position of the axis of a cylindrically symmetric flux tube. From hadron spectroscopy, the string constant, i.e. the amount of energy per unit length, is deduced to be $ \kappa \approx 1$ GeV/fm. The expression `massless' relativistic string is somewhat of a misnomer: $\kappa$ effectively corresponds to a `mass density' along the string. Let us now turn to the fragmentation process. As the $\mathrm{q}$ and $\overline{\mathrm{q}}$ move apart, the potential energy stored in the string increases, and the string may break by the production of a new $\mathrm{q}' \overline{\mathrm{q}}'$ pair, so that the system splits into two colour-singlet systems $\mathrm{q} \overline{\mathrm{q}}'$ and $\mathrm{q}' \overline{\mathrm{q}}$. If the invariant mass of either of these string pieces is large enough, further breaks may occur. In the Lund string model, the string break-up process is assumed to proceed until only on-mass-shell hadrons remain, each hadron corresponding to a small piece of string with a quark in one end and an antiquark in the other. In order to generate the quark--antiquark pairs $\mathrm{q}' \overline{\mathrm{q}}'$ which lead to string break-ups, the Lund model invokes the idea of quantum mechanical tunnelling. This leads to a flavour-independent Gaussian spectrum for the $p_{\perp}$ of $\mathrm{q}' \overline{\mathrm{q}}'$ pairs. Since the string is assumed to have no transverse excitations, this $p_{\perp}$ is locally compensated between the quark and the antiquark of the pair. The total $p_{\perp}$ of a hadron is made up out of the $p_{\perp}$ contributions from the quark and antiquark that together form the hadron. Some contribution of very soft perturbative gluon emission may also effectively be included in this description. The tunnelling picture also implies a suppression of heavy-quark production, $\u : \d : \mathrm{s} : \c \approx 1 : 1 : 0.3 : 10^{-11}$. Charm and heavier quarks hence are not expected to be produced in the soft fragmentation, but only in perturbative parton-shower branchings $\mathrm{g} \to \mathrm{q} \overline{\mathrm{q}}$. When the quark and antiquark from two adjacent string breakings are combined to form a meson, it is necessary to invoke an algorithm to choose between the different allowed possibilities, notably between pseudoscalar and vector mesons. Here the string model is not particularly predictive. Qualitatively one expects a $1 : 3$ ratio, from counting the number of spin states, multiplied by some wave-function normalization factor, which should disfavour heavier states. A tunnelling mechanism can also be used to explain the production of baryons. This is still a poorly understood area. In the simplest possible approach, a diquark in a colour antitriplet state is just treated like an ordinary antiquark, such that a string can break either by quark--antiquark or antidiquark--diquark pair production. A more complex scenario is the `popcorn' one, where diquarks as such do not exist, but rather quark--antiquark pairs are produced one after the other. This latter picture gives a less strong correlation in flavour and momentum space between the baryon and the antibaryon of a pair. In general, the different string breaks are causally disconnected. This means that it is possible to describe the breaks in any convenient order, e.g. from the quark end inwards. One therefore is led to write down an iterative scheme for the fragmentation, as follows. Assume an initial quark $\mathrm{q}$ moving out along the $+z$ axis, with the antiquark going out in the opposite direction. By the production of a $\mathrm{q}_1 \overline{\mathrm{q}}_1$ pair, a meson $\mathrm{q} \overline{\mathrm{q}}_1$ is produced, leaving behind an unpaired quark $\mathrm{q}_1$. A second pair $\mathrm{q}_2 \overline{\mathrm{q}}_2$ may now be produced, to give a new meson $\mathrm{q}_1 \overline{\mathrm{q}}_2$, etc. At each step the produced hadron takes some fraction of the available energy and momentum. This process may be iterated until all energy is used up, with some modifications close to the $\overline{\mathrm{q}}$ end of the string in order to make total energy and momentum come out right. The choice of starting the fragmentation from the quark end is arbitrary, however. A fragmentation process described in terms of starting at the $\overline{\mathrm{q}}$ end of the system and fragmenting towards the $\mathrm{q}$ end should be equivalent. This `left--right' symmetry constrains the allowed shape of the fragmentation function $f(z)$, where $z$ is the fraction of the remaining light-cone momentum $E \pm p_z$ (+ for the $\mathrm{q}$ jet, $-$ for the $\overline{\mathrm{q}}$ one) taken by each new particle. The resulting `Lund symmetric fragmentation function' has two free parameters, which are determined from data. If several partons are moving apart from a common origin, the details of the string drawing become more complicated. For a $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ event, a string is stretched from the $\mathrm{q}$ end via the $\mathrm{g}$ to the $\overline{\mathrm{q}}$ end, i.e. the gluon is a kink on the string, carrying energy and momentum. As a consequence, the gluon has two string pieces attached, and the ratio of gluon to quark string force is 2, a number which can be compared with the ratio of colour charge Casimir operators, $N_C/C_F = 2/(1-1/N_C^2) = 9/4$. In this, as in other respects, the string model can be viewed as a variant of QCD where the number of colours $N_C$ is not 3 but infinite. Note that the factor 2 above does not depend on the kinematical configuration: a smaller opening angle between two partons corresponds to a smaller string length drawn out per unit time, but also to an increased transverse velocity of the string piece, which gives an exactly compensating boost factor in the energy density per unit string length. The $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ string will fragment along its length. To first approximation this means that there is one fragmenting string piece between $\mathrm{q}$ and $\mathrm{g}$ and a second one between $\mathrm{g}$ and $\overline{\mathrm{q}}$. One hadron is straddling both string pieces, i.e. sitting around the gluon corner. The rest of the particles are produced as in two simple $\mathrm{q} \overline{\mathrm{q}}$ strings, but strings boosted with respect to the overall c.m. frame. When considered in detail, the string motion and fragmentation is more complicated, with the appearance of additional string regions during the time evolution of the system. These corrections are especially important for soft and collinear gluons, since they provide a smooth transition between events where such radiation took place and events where it did not. Therefore the string fragmentation scheme is `infrared safe' with respect to soft or collinear gluon emission. For events that involve many partons, there may be several possible topologies for their ordering along the string. An example would be a $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}_1 \mathrm{g}_2$ (the gluon indices are here used to label two different gluon-momentum vectors), where the string can connect the partons in either of the sequences $\mathrm{q} - \mathrm{g}_1 - \mathrm{g}_2 - \overline{\mathrm{q}}$ and $\mathrm{q} - \mathrm{g}_2 - \mathrm{g}_1 - \overline{\mathrm{q}}$. The matrix elements that are calculable in perturbation theory contain interference terms between these two possibilities, which means that the colour flow is not always well-defined. Fortunately, the interference terms are down in magnitude by a factor $1/N_C^2$, where $N_C = 3$ is the number of colours, so approximate recipes can be found. In the leading log shower description, on the other hand, the rules for the colour flow are well-defined. A final comment: in the argumentation for the importance of colour flows there is a tacit assumption that soft-gluon exchanges between partons will not normally mess up the original colour assignment; this is likely the case but has not been proven. \subsection{Decays} A large fraction of the particles produced by fragmentation are unstable and subsequently decay into the observable stable (or almost stable) ones. It is therefore important to include all particles with their proper mass distributions and decay properties. Although involving little deep physics, this is less trivial than it may sound: while a lot of experimental information is available, there is also very much that is missing. For charm mesons, it is necessary to put together measured exclusive branching ratios with some inclusive multiplicity distributions to obtain a consistent and reasonably complete set of decay channels, a rather delicate task. For bottom, so far only a rather simple phase-space type of generator has been used for hadronic decays. Normally it is assumed that decay products are distributed according to phase space, i.e. that there is no dynamics involved in their relative distribution. However, in many cases additional assumptions are necessary, e.g. for semileptonic decays of charm and bottom hadrons one needs to include the proper weak matrix elements. Particles may also be produced polarized and impart a non-isotropic distribution to their decay products. Many of these effects are not at all treated in the program. In fact, spin information is not at all carried along, but has to be reconstructed explicitly when needed. The normal decay treatment is handled by {\tsc{Jetset}}, making use of a set of tables where branching ratios and decay modes are stored. In {\tsc{Pythia}} a separate decay treatment exists, used exclusively for a specific list of particles: $\mathrm{Z}^0$, $\mathrm{W}^{\pm}$, $\H^0$, $\mathrm{Z}'^0$, $\mathrm{W}'^{\pm}$, $\H'^0$, $\mathrm{A}^0$, $\H^{\pm}$, $\eta_{\mrm{tech}}^0$, $\mathrm{R}^0$, $\mathrm{q}^*$, $\ell^*$, and the leptoquark $\L_{\mathrm{Q}}$. Together we call these resonances, and contrast the `particle decay' treatment of {\tsc{Jetset}} with the `resonance decay' one of {\tsc{Pythia}}. Of course, this is just a matter of terminology: a particle like the $\rho$ could also be called a resonance. What characterizes a ({\tsc{Pythia}}) resonance is that partial widths and branching ratios are calculated dynamically, as a function of the actual mass of a particle. Therefore not only do branching ratios change between an $\H^0$ of nominal mass 100 GeV and one of 200 GeV, but also for a Higgs of nominal mass 200 GeV, the branching ratios would change between an actual mass of 190 GeV and 210 GeV, say. This is particularly relevant for reasonably broad resonances, and in threshold regions. For an approach like this to work, it is clearly necessary to have perturbative expressions available for all partial widths, which is one reason why a corresponding treatment would not be the same for an ordinary hadronic resonance, like the $\rho$. The decay products of {\tsc{Pythia}} resonances are typically quarks, leptons, or other resonances, e.g. $\mathrm{W} \to \mathrm{q} \overline{\mathrm{q}}'$ or $\H^0 \to \mathrm{W}^+ \mathrm{W}^-$. In decays to quarks, parton showers are automatically added to give a more realistic multijet structure, and one may also allow photon emission off leptons. If the decay products in turn are resonances, further decays are necessary. Often spin information is available in resonance decay matrix elements, contrary to the normal state of affairs in ordinary particle decays. This means that the angular orientations in the two decays of a $\mathrm{W}^+ \mathrm{W}^-$ pair are properly correlated. Occasionally, the information is not available, and then resonances decay isotropically. The top quark is a special problem. The original machinery is based on the assumption that the $\t$ is long-lived, so that top hadrons have time to form in the fragmentation process, and afterwards these mesons decay weakly. With current `best bet' mass values, this is not correct, but one should rather consider top decay before fragmentation. Top should then be handled like one of the above resonances. Therefore the program now contains an alternative along these lines, which is the preferred option. \clearpage \section{Program Overview} This section contains a diverse collection of information. The first part is an overview of previous {\tsc{Jetset}} and {\tsc{Pythia}} versions. The second gives instructions for installation of the programs and describes their philosophy: how they are constructed and how they are supposed to be used. It also contains some information on how to read this manual. The third and final part contains several examples of pieces of code or short programs, to illustrate the general style of program usage. The last part is mainly intended as an introduction for completely new users, and can be skipped by more experienced ones. Since the {\tsc{Jetset}} and {\tsc{Pythia}} programs today are so closely connected, and are gradually coalescing, they are presented together in this report. However, they still appear as separate entities, with slightly different style and emphasis. {\tsc{Jetset}} is the older of the two, and is at the origin of the whole `Lund' family of event generators. It can be subdivided in two parts. The larger is a generic package for jet fragmentation, particle decays, final-state parton showers, event-analysis routines, and other utitilies. This package can be used in the context of any hard process, provided one is willing to buy the underlying assumption of jet universality, i.e. that the fragmentation process is fundamentally the same whether one is considering an $\e^+\e^-$ or a $\p\p$ event, and that the only differences are to be found in the parton-level processes involved. This package is not only used by all other `Lund' programs, but also by numerous other programs written to study specific processes. The smaller part of {\tsc{Jetset}} is a generator for $\e^+\e^-$ annihilation events, according to either a parton-shower or a matrix-element approach. The {\tsc{Jetset}} program is completely selfcontained. {\tsc{Pythia}} is a program made to generate hard or soft processes in collisions between leptons, hadrons and photons, especially at $\e^+\e^-$, $\e\p$ and $\p\p$ colliders. Where {\tsc{Jetset}} is a loose collection of routines that you can combine as desired, {\tsc{Pythia}} is a more structured program, where you initially set up what processes you want to study, and thereafter all events will be generated according to this specification. Included is an extensive library of hard subprocess differential cross sections, a library of parton distributions, a process generation machinery, treatment of initial-state showers and beam remnants, and a few odds and ends. {\tsc{Jetset}} is used for final-state showers, fragmentation and decay, but no other external libraries are needed. An interface to external parton-distribution function libraries is provided, however. Many programs written by other persons make use of {\tsc{Jetset}}, and a few also of {\tsc{Pythia}}. It is not my intention to give a complete list here. A majority of these programs are specific to given collaborations, and therefore not publicly distributed. Below we give a list of a few public programs from the `Lund group', which may have a somewhat wider application. None of them are supported by the current author, so any requests should be directed to the persons mentioned. \begin{Itemize} \item \tsc{Ariadne} is a generator for dipole emission, written mainly by L. L\"onnblad \cite{Pet88}. The dipole provides an alternative formulation of initial- and final-state showers. {\tsc{Jetset}} or {\tsc{Pythia}} can be used to generate the hard process and {\tsc{Jetset}} to do the fragmentation. \item \tsc{Aroma} is a generator for heavy-flavour processes in leptoproduction, written by G.~Ingelman and G. Schuler \cite{Ing88}. It uses {\tsc{Jetset}} for fragmentation. \item \tsc{Fritiof} is a generator for hadron--hadron, hadron--nucleus and nucleus--nucleus collisions \cite{Nil87}, which makes use of {\tsc{Pythia}} to generate hard QCD scatterings and of {\tsc{Jetset}} for fragmentation. Currently H. Pi is responsible for program development. \item \tsc{Lepto} is a leptoproduction event generator, written mainly by G. Ingelman \cite{Ing80}. It can generate parton configurations in deep inelastic scattering according to a number of possibilities. It makes use of {\tsc{Jetset}} for fragmentation and additionally has a parton-shower option based on {\tsc{Pythia}}. \item \tsc{Lucifer} is a photoproduction generator written by G. Ingelman and A. Weigend \cite{Ing87a}. It is a modification of an earlier version of {\tsc{Pythia}} and makes use of {\tsc{Jetset}}. \item \tsc{Pompyt} is a generator for pomeron interactions written by P. Bruni and G. Ingelman \cite{Bru93}. This program defines parton distributions, flux factors and other aspects specific to the pomeron, which is combined with the standard {\tsc{Pythia}} machinery for process generation. \item \tsc{Twister} is a generator for higher-twist processes, written by G. Ingelman \cite{Ing87}. It is a modification of an earlier version of {\tsc{Pythia}} and makes use of {\tsc{Jetset}}. \end{Itemize} One should also note that a version of {\tsc{Pythia}} has been modified to include the effects of longitudinally polarized incoming protons. This is the work of St. G\"ullenstern et al. \cite{Gul93}. \subsection{Update History} Both {\tsc{Jetset}} and {\tsc{Pythia}} are by now fairly old and well-established programs, but they are still steadily being improved on. While evolution was especially rapid for {\tsc{Jetset}} in the early days, that program has by now reached a certain level of maturity, and the pace of change has dropped significantly. {\tsc{Pythia}}, on the other hand, has been continually extended in recent years, and may still see further growth, although most of the basic structure should be in place by now. In earlier days, before the advent of electronic mail, programs were only infrequently distributed, and version numbers corresponded to distinct new upgrades. Today, the evolutionary process is more continuous and so is the distribution of new versions. In particular, the introduction of a new process or feature is often done on short notice, if no problems of backwards compatibility are involved. With this distribution, the subversion numbers have therefore been expanded to three digits, where the last two give sub-subversions. For every change made in the public file, the sub-subversion number is updated, together with the `last date of change'. In most referencing the shorter `{\tsc{Jetset}} version 7.4' could still be preferable to e.g. `{\tsc{Jetset}} version 7.412'. For the record, in Tables \ref{Jetsetver} and \ref{Pythiaver} we list the official main versions of {\tsc{Jetset}} and {\tsc{Pythia}}, respectively, with some brief comments. \begin{table}[tp] \captive{The main versions of {\tsc{Jetset}}, with their date of appearance, published manuals, and main changes from previous versions. \label{Jetsetver}} \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c|l|@{\protect\rule{0mm}{\tablinsep}}} \hline No. & Date & Publ. & Main new or improved features \\[1mm] \hline 1 & Nov 78 & \cite{Sjo78} & single-quark jets \\ 2 & May 79 & \cite{Sjo79} & heavy-flavour jets \\ 3.1 & Aug 79 & --- & 2-jets in $\e^+\e^-$, preliminary 3-jets \\ 3.2 & Apr 80 & \cite{Sjo80} & 3-jets in $\e^+\e^-$ with full matrix elements, \\ & & & toponium $\to \mathrm{g} \mathrm{g} \mathrm{g}$ decays \\ 3.3 & Aug 80 & --- & softer fragmentation spectrum \\ 4.1 & Apr 81 & --- & baryon production and diquark fragmentation, \\ & & & fourth-generation quarks, larger jet systems \\ 4.2 & Nov 81 & --- & low-$p_{\perp}$ physics \\ 4.3 & Mar 82 & \cite{Sjo82} & 4-jets and QFD structure in $\e^+\e^-$, \\ & Jul 82 & \cite{Sjo83} & event-analysis routines \\ 5.1 & Apr 83 & --- & improved string fragmentation scheme, symmetric \\ & & & fragmentation, full 2$^{\mrm{nd}}$ order QCD for $\e^+\e^-$ \\ 5.2 & Nov 83 & --- & momentum-conservation schemes for IF, \\ & & & initial-state photon radiation in $\e^+\e^-$ \\ 5.3 & May 84 & --- & `popcorn' model for baryon production \\ 6.1 & Jan 85 & --- & common blocks restructured, parton showers \\ 6.2 & Oct 85 & \cite{Sjo86} & error detection \\ 6.3 & Oct 86 & \cite{Sjo87} & new parton-shower scheme \\ 7.1 & Feb 89 & --- & new particle codes and common block structure, \\ & & & more mesons, improved decays, vertex information, \\ & & & Abelian gluon model, Bose--Einstein effects \\ 7.2 & Nov 89 & --- & interface to new standard common block, \\ & & & photon emission in showers \\ 7.3 & May 90 & \cite{Sjo92d} & expanded support for non-standard particles \\ 7.4 & Dec 93 & \cite{Sjo94} & updated particle data and defaults \\[1mm] \hline \end{tabular} \end{center} \end{table} \begin{table}[tp] \captive{The main versions of {\tsc{Pythia}}, with their date of appearance, published manuals, and main changes from previous versions. \label{Pythiaver}} \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c|l|@{\protect\rule{0mm}{\tablinsep}}} \hline No. & Date & Publ. & Main new or improved features \\[1mm] \hline 1 & Dec 82 & \cite{Ben84} & synthesis of predecessors \tsc{Compton}, \tsc{Highpt} and \\ & & & \tsc{Kassandra} \\ 2 & --- & & \\ 3.1 & --- & & \\ 3.2 & --- & & \\ 3.3 & Feb 84 & \cite{Ben84a} & scale-breaking parton distributions \\ 3.4 & Sep 84 & \cite{Ben85} & more efficient kinematics selection \\ 4.1 & Dec 84 & & initial- and final-state parton showers, $\mathrm{W}$ and $\mathrm{Z}$ \\ 4.2 & Jun 85 & & multiple interactions \\ 4.3 & Aug 85 & & $\mathrm{W} \mathrm{W}$, $\mathrm{W} \mathrm{Z}$, $\mathrm{Z} \mathrm{Z}$ and $\mathrm{R}$ processes \\ 4.4 & Nov 85 & & $\gamma \mathrm{W}$, $\gamma \mathrm{Z}$, $\gamma \gamma$ processes \\ 4.5 & Jan 86 & & $\H^0$ production, diffractive and elastic events \\ 4.6 & May 86 & & angular correlation in resonance pair decays \\ 4.7 & May 86 & & $\mathrm{Z}'^0$ and $\H^+$ processes \\ 4.8 & Jan 87 & \cite{Ben87} & variable impact parameter in multiple interactions \\ 4.9 & May 87 & & $\mathrm{g} \H^+$ process \\ 5.1 & May 87 & & massive matrix elements for heavy quarks \\ 5.2 & Jun 87 & & intermediate boson scattering \\ 5.3 & Oct 89 & & new particle and subprocess codes, new common block \\ & & & structure, new kinematics selection, some \\ & & & lepton--lepton and lepton--hadron interactions, \\ & & & new subprocesses \\ 5.4 & Jun 90 & & $s$-dependent widths, resonances not on the mass shell, \\ & & & new processes, new parton distributions \\ 5.5 & Jan 91 & & improved $\e^+\e^-$ and $\e\p$, several new processes \\ 5.6 & Sep 91 & \cite{Sjo92d} & reorganized parton distributions, new processes, \\ & & & user-defined external processes \\ 5.7 & Dec 93 & \cite{Sjo94} & new total cross sections, photoproduction, top decay \\[1mm] \hline \end{tabular} \end{center} \end{table} All versions preceding {\tsc{Jetset}} 7.3 and {\tsc{Pythia}} 5.6 should now be considered obsolete, and are no longer supported. For stable applications, the earlier combination {\tsc{Jetset}} 6.3 and {\tsc{Pythia}} 4.8 could still be used, however. {\tsc{Jetset}} version 7 and {\tsc{Pythia}} version 5 have been evolved in parallel, so some of the processes added in later versions of {\tsc{Pythia}} make use of particle data only found in {\tsc{Jetset}} from that time onwards. Although it would be possible to combine {\tsc{Pythia}} 5.7 with {\tsc{Jetset}} 7.3, e.g., it is not recommended. From the current versions onwards, checks have therefore been introduced to detect the use of (potentially) incompatible subversions, with warnings issued at initialization if that should be the case. Previous versions of the manuals have contained detailed lists of modifications from one version to the next, see e.g. \cite{Sjo92d}. Below we only reproduce the updates that appear with the most recent versions of the programs. Some of them were introduced in later editions of {\tsc{Pythia}} 5.6 with {\tsc{Jetset}} 7.3, while others are completely new. If nothing is explicitly said, these changes do not affect backwards compatibility, but only add new features. \subsubsection{Updates in JETSET 7.4} Changes from version 7.3 to 7.4 are not so large, although the impact of the updated particle data and parameter default values may need to be studied. \begin{Itemize} \item Particle data have been updated in accordance with the 1992 Review of Particle Properites \cite{PDG92}. (As usual, with a free interpretation of inconsistencies, unclarities and other gaps in the knowledge.) Changes are especially drastic for charm and bottom. In the bottom sector the decay properties are now given individually for $\mathrm{B}^0$, $\mathrm{B}^+$, $\mathrm{B}_{\mathrm{s}}^0$, $\mathrm{B}_{\c}^+$ and $\Lambda_{\b}^0$, i.e. the generic data for `pseudoparticle' 85 are only used for other weakly decaying $\mathrm{B}$ baryons. \item Also a few other Standard Model parameters have been updated, such as the $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$ masses and widths, $\sin^2 \! \theta_W$ and the CKM matrix elements. \item Fragmentation and parton shower parameters have been modified to reflect current LEP knowledge \cite{LEP90}, i.e. a minor retuning starting from an average of the `best' parameter values obtained by the four LEP collaborations. Bose-Einstein effects are still left out. Flavour composition is unchanged, except for a suppression of $\eta'$ production. Affected by the change are \ttt{MSTJ(11)}, \ttt{PARJ(21)}, \ttt{PARJ(23)}, \ttt{PARJ(24)}, \ttt{PARJ(41)}, \ttt{PARJ(42)}, \ttt{PARJ(54)}--\ttt{PARJ(58)} and \ttt{PARJ(81)}. \item Several other default values have been changed for switches and parameters in the $\e^+\e^-$, parton shower and fragmentation parts of the programs. These changes are intended to reflect our current best knowledge. See \ttt{MSTJ(26)}, \ttt{MSTJ(27)}, \ttt{MSTJ(41)}, \ttt{MSTJ(46)}, \ttt{MSTJ(50)}, \ttt{MSTJ(110)}, \ttt{PARJ(26)}, and \ttt{PARJ(121)}--\ttt{PARJ(125)}. \item A common title page for {\tsc{Jetset}} and {\tsc{Pythia}} has been introduced with the \ttt{LULOGO} routine. Sub-subversion numbers are also given. \item Several options have been added for the \ttt{LUSHOW} shower routine. See \ttt{MSTJ(41)}, \ttt{MSTJ(47)} and \ttt{MSTJ(50)}. \item A $\b$ quark produced in the decay of a top hadron is allowed to radiate according to the standard parton shower scheme. \item The scalar gluon option contains the full electroweak angular distribution of 3-jet events. \item The \ttt{LUCOMP} routine has been modified. Among other things, the $\mathrm{B}_{\c}^+$ now appears as a separate compressed code, further codes for diffractive states have been added to the current list, and the pomeron (reggeon, $\eta_{\mrm{techni}}$) has been added as particle 29 (28, 38). \item A minimum threshold for calorimeter cell energy has been introduced for the \ttt{LUCELL} routine. \item All obsolescent features of the Fortran 90 standard have been removed, i.e. the program should work well either with a Fortran 77 compiler or with a Fortran 90 one. \item A few minor errors have been corrected. \end{Itemize} The following changes have been made since the beginning of 1994, i.e. since the original distribution 7.400: \begin{Enumerate} \item {\tsc{Jetset}} version 7.401, 11 February 1994: \begin{Itemize} \item Protect against overflow in \ttt{LUZDIS} (needed on some machines). \end{Itemize} \item {\tsc{Jetset}} version 7.402, 7 April 1994: \begin{Itemize} \item New option to suppress either hard or soft radiation in \ttt{LUSHOW}, see \ttt{MSTJ(40)}. \item A generic interface to an external $\tau$ decay library has been introduced, see \ttt{MSTJ(28)} and \ttt{SUBROUTINE LUTAUD}. \item In a few places, a dot has been moved from the end of one line to the beginning of the next continuation line, or the other way around, to keep together tokens such as \ttt{.EQ.} or \ttt{.AND.}, since some debuggers may otherwise complain. \item A source of (harmless) division by zero in \ttt{LUSHOW} has been removed. \end{Itemize} \item {\tsc{Jetset}} version 7.403, 15 July 1994: \begin{Itemize} \item Leptons and photons which are unrelated to the system feeling the Bose-Einstein effects do not have their energies and momenta changed in the global rescaling step of \ttt{LUBOEI}. (Example: $\mathrm{W}^+\mathrm{W}^-$ events, where one $\mathrm{W}$ decays leptonically; before these lepton momenta could be slightly changed, but now not.) Further, the \ttt{LUBOEI} routine has been changed to avoid an unintentional gap in the limits of the very first bin. \item The option \ttt{LUEDIT(16)} (used e.g. from \ttt{PYEVNT}) has been improved with a more extensive search for missing daughter pointers. \item The \ttt{KLU(I,16)} procedure for finding rank has been rewritten to work in the current {\tsc{Jetset}} version, which it did not before. However, note that it will only work for \ttt{MSTU(16)=2}. As a general comment, the options 14--17 of \ttt{KLU} were written at a time when possible event histories were less complex, and can not be guaranteed always to work today. \end{Itemize} \item {\tsc{Jetset}} version 7.404, 25 August 1994: \begin{Itemize} \item \ttt{LUSHOW} has been corrected, so that if $\t$, $\mathrm{l}$ or $\mathrm{h}$ quarks (or $\d^*$ or $\u^*$ quarks masked as $\mathrm{l}$ or $\mathrm{h}$ ones) are given with masses that vary from event to event (a Breit-Wigner shape, e.g.), the current mass rather than the nominal mass is used to define the cut-off scales of parton shower evolution. \item \ttt{LULOGO} has been modified to take into account that a new {\tsc{Pythia/Jetset}} description has been published in \\ T. Sj\"ostrand, Computer Phys. Commun. {\bf 82} (1994) 74 \\ and is from now on the standard reference to these two programs. \end{Itemize} \item {\tsc{Jetset}} version 7.405, 27 January 1995: \begin{Itemize} \item \ttt{LUCELL} has been corrected, in that in the option with smearing of energy rather than transverse energy, the conversion factor between the two was applied in the wrong direction. \item \ttt{LUSHOW} has been corrected in one place where the PMTH array was addressed with the wrong order of the indices. This affected quark mass corrections in the matching to the three-jet matrix elements. \item An additional check has been included in \ttt{LUBOEI} that there are at least two particles involved in the Bose-Einstein effects. (No problem except in some bizarre situations.) \end{Itemize} \item {\tsc{Jetset}} version 7.406, 20 February 1995: \begin{Itemize} \item A new option has been added for the behaviour of the running $\alpha_{\mathrm{em}}(Q^2)$ in \ttt{ULALEM}. This is not added as a true physics scenario, but only to produce results with a given, fixed value for the hard events, while still keeping the conventional value in the $Q^2=0$ limit. See \ttt{MSTU(101)}, \ttt{PARU(103)}, \ttt{PARU(104)}. Additionally, the $G_{\mrm{F}}$ constant has been added to the parameter list, see \ttt{PARU(105)}. \item The LULOGO routine has been updated to reflect my change of affiliation. \end{Itemize} \item {\tsc{Jetset}} version 7.407, 21 June 1995: \begin{Itemize} \item Header and \ttt{LULOGO} have been updated with respect to phone number and WWW access. \item The \ttt{PHEP} and \ttt{VHEP} variables in the \ttt{/HEPEVT/} common block are now assumed to be in \ttt{DOUBLE PRECISION}, in accord with the proposed LEP 2 workshop addendum to the standard. \item In \ttt{LUTEST} a missing decimal point on the energy check has been reinstated. \item In \ttt{LUINDF} an expression has been protected against vanishing denominator. \end{Itemize} \item {\tsc{Jetset}} version 7.408, 23 August 1995: \begin{Itemize} \item Check against division by zero in \ttt{LUSHOW}. \end{Itemize} \end{Enumerate} \subsubsection{Updates in PYTHIA 5.7} The updates from version 5.6 to 5.7 are all minor, and just about any program that ran with version 5.6 will also work with {\tsc{Pythia}} 5.7. However, as for {\tsc{Jetset}}, it should be noted that some important default values have been changed. \begin{Itemize} \item New parametrizations of the total cross sections of hadronic reactions, based in Donnachie--Landshoff \cite{Don92}, which replace the old ones. \item New parametrizations of elastic and single and double diffractive cross sections of hadronic reactions, based on Schuler--Sj\"ostrand \cite{Sch94,Sch93a}, which replace the old ones. Also the slope parameters, the diffractive mass distributions and other aspects of the event generation have been changed accordingly. \item A possibility to give own total, elastic and diffractive cross sections. \item The single diffractive cross section has been split into its two constituents, $AB \to XB$ and $AB \to AX$. As a consequence, the diffractive subprocess codes 92--94 have received changed meaning. \item A new common block \ttt{PYINT7} has been added for the expanded total cross section information, and this information has been partly removed from other common blocks. \item A much extended description of photoproduction physics, with the possibility to simulate separately VMD, anomalous and direct processes \cite{Sch93,Sch93a}. \item The selection of proton parton distributions that come with the program has been updated with the CTEQ2 ones, while some others have been removed. New default is the leading-order fit CTEQ2L. \item Since the \tsc{Pdflib} library now has been expanded to contain also parton-distribution functions for the photon, the interfaces to the \tsc{Pakpdf} and \tsc{Phopdf} libraries have been removed. In addition, the interface to \tsc{Pdflib} has been modified, and is now for appropriate for \tsc{Pdflib} version 4. \item An extension of hadron parton distributions into the low-$x$ and low-$Q^2$ region \cite{Sch93a}. \item The top quark can be made to decay before it has time to fragment. In view of the current best estimate for the top mass, this is the expected behaviour, and is therefore now default. Further, a parton shower is allowed to evolve in the top decay. Also fourth generation quarks are allowed to decay before they fragment, and so on. \item It is possible to call \ttt{PYEVNT} with energies that vary from one event to the next, without the need to reinitialize. \item Improved scheme for post-factor conservation of $x$ and $Q^2$ in deep inelastic scattering. \item Processes 15, 19, 30 and 35 have been expanded to cover $\gamma^*$ production in addition to the $\mathrm{Z}^0$ one, with full interference. \item New process 80, $\mathrm{q}\gamma \to \mathrm{q}'\pi^{\pm}$. \item New process 110, $\mathrm{f}\overline{\mathrm{f}} \to \gamma\H^0$. \item New process 149, $\mathrm{g}\g \to \eta_{\mrm{techni}}$. \item New option for initial state radiation to restrict angular range of emission in accordance with coherence considerations. \item Some options have been added or removed, and default values have been changed. This includes \ttt{KFIN} (top parton distributions off by default), \ttt{MSTP(7)}, \ttt{MSTP(11)}, \ttt{MSTP(14)}, \ttt{MSTP(23)}, \ttt{MSTP(30)} (removed), \ttt{MSTP(31)}, ~\ttt{MSTP(34)}, \ttt{MSTP(45)}, \ttt{MSTP(48)}, \ttt{MSTP(49)}, \ttt{MSTP(62)}, \ttt{MSTP(67)}, \ttt{MSTP(101)}, \ttt{PARP(13)}, \ttt{PARP(81)}, \ttt{PARP(82)}, \ttt{PARP(47)} and \ttt{PARP(101)}. \item All obsolescent features of the Fortran 90 standard have been removed, i.e. the program should work well either with a Fortran 77 compiler or with a Fortran 90 one. \item A few minor errors have been corrected. \end{Itemize} The following changes have been made since the beginning of 1994, i.e. since the original distribution 5.700: \begin{Enumerate} \item {\tsc{Pythia}} version 5.701, 27 January 1994: \begin{Itemize} \item The machinery to handle $\gamma\gamma$ interactions is expanded to the level already available for $\gamma\mathrm{p}$. This in particular means that a number of new options appear for \ttt{MSTP(14)}. Affected are also \ttt{MINT(105)}, \ttt{MINT(107)}, \ttt{MINT(108)}, \ttt{MINT(109)}, \ttt{VINT(282)} (removed), \ttt{VINT(283)} and \ttt{VINT(284)}. Parametrizations are introduced for meson--meson total, elastic and diffractive cross sections, needed for the VMD part of the photon. The treatment of cross sections for hard processes, of initial state radiation, of beam remnants and of other aspects are also expanded to cover the new possibilities. A first study of the relevant physics aspects is found in \cite{Sch94a}. \item An option is introduced to modify the $Q^2$ scale of the anomalous part of the photon parton distributions, see \ttt{MSTP(59)} and \ttt{PARP(59)}. \item Correction of an error, where the generation of jet and low-$p_{\perp}$ events could give incorrect cross section information with \ttt{PYSTAT(1)} at low energies. The event generation itself was correct. (The error was introduced as a consequence of allowing variable energies.) \item A rejection is introduced for top events where the top mass (selected according to a Breit-Wigner) is too low to allow the decay into a $\mathrm{W}$ on the mass shell. \item The correction of a few other minor bugs, probably harmless. \end{Itemize} \item {\tsc{Pythia}} version 5.702, 13 February 1994: \begin{Itemize} \item The interface to \tsc{Pdflib} has been modified to reflect that \ttt{TMAS} should no longer be set except in first \ttt{PDFSET} call. (Else a huge amount of irrelevant warning messages are generated by \tsc{Pdflib}.) \item The \ttt{STOP} statement in a few dummy routines has been modifed to avoid irrelevant compilation warning messages on IBM mainframes. \item A few labels have been renumbered. \end{Itemize} \item {\tsc{Pythia}} version 5.703, 22 February 1994: \begin{Itemize} \item Removal of a bug in \ttt{PYRESD}, which could give (under some specific conditions) errors in the colour flow. \end{Itemize} \item {\tsc{Pythia}} version 5.704, 7 April 1994: \begin{Itemize} \item Process 11 has been corrected, for the part that concerns anomalous couplings (contact interactions) in the $\mathrm{q}\q' \to \mathrm{q}\q'$ process. The error was present in the expression for $\u\overline{\mathrm{d}} \to \u\overline{\mathrm{d}}$ and obvious permutations, while $\u\d \to \u\d$, $\u\overline{\mathrm{u}} \to \u\overline{\mathrm{u}}$ and the others were correct. \item The option with post-facto $(x,Q^2)$ conservation in deep inelastic scattering can give infinite loops when applied to process 83, in particular if one asks for the production of a top. (Remember that the standard DIS kinematics is defined for massless quarks.) Therefore the switch \ttt{MSTP(23)} has been modifed so that by default only process 10 is affected. \item \ttt{PYRESD} is modified to ensure isotropic angular distributions in the decays of the top or a fourth generation particle, i.e. in $\t \to \b\mathrm{W}^+$. This may not be the correct distribution but, unless explicit knowledge exists for a given process, this should always be the default. \item In processes 16, 20, 31 and 36 the $\mathrm{W}$ propagator has been modified to include $s$-dependent widths in the Breit-Wigner shape. The most notable effect is a suppression of the low-mass tail of the $\mathrm{W}$ mass spectrum. \item When \tsc{Pdflib} is used, \ttt{PDFSET} is now only called whenever a different structure function is requested. For $\mathrm{p}\p$ events therefore only one call is made, while $\gamma\mathrm{p}$ interactions still involve a call to \ttt{PDFSET} for each \ttt{STRUCTM} one, since $\gamma$ and $\mathrm{p}$ structure functions have to be called alternatingly. \ttt{MINT(93)} is used to keep track of latest structure function called. \item In a few places, a dot has been moved from the end of one line to the beginning of the next continuation line, or the other way around, to keep together tokens such as \ttt{.EQ.} or \ttt{.AND.}, since some debuggers may otherwise complain. \item A number of minor errors have been corrected. \end{Itemize} \item {\tsc{Pythia}} version 5.705, 15 July 1994: \begin{Itemize} \item A completely new possibility to have {\tsc{Pythia}} mix different allowed processes (direct, VMD and anomalous) in $\gamma\mathrm{p}$ and $\gamma\gamma$ interactions. This option can be accessed with \ttt{MSTP(14)=10}. The relevant physics description and programming details may be found in sections \ref{sss:photoprod} and \ref{sss:photoprodclass}. This facility is still not definitive, in that it is hoped to gradually enhance it with further features. The cross-section output of the \ttt{PYSTAT} has been expanded to reflect the further subdivision of the total cross section. \item The new facility above has required a major restructuring of some of the code: the routine \ttt{PYEVKI} has been removed, new routines \ttt{PYINBM}, \ttt{PYINPR} and \ttt{PYSAVE} created, and some material has been moved to or from \ttt{PYINIT}, \ttt{PYINRE} and \ttt{PYINKI}. New variables include \ttt{MSTI(9)}, \ttt{MINT(121)}, \ttt{MINT(122)}, \ttt{MINT(123)} and \ttt{VINT(285)}. \item The GRV leading-order dynamically generated parton distributions for the $\mathrm{p}$ and $\pi$ have been included as options, see \ttt{MSTP(51)} and \ttt{MSTP(53)}. \item A parametrization of the homogeneous solution to the anomalous photon parton distributions have been added as an option, see \ttt{MSTP(56)=3}. \item The treatment of the anomalous photon component can be modified with the new switch \ttt{MSTP(15)} and variable \ttt{PARP(17)}; at the same time \ttt{MSTP(59)} and \ttt{PARP(59)} have been removed. The new options are mainly intended for comparative studies and should not normally be touched. \item The option \ttt{MSTP(92)=5} for beam remnant treatment erroneously missed some statements which now have been inserted. Further, new options have been added for the beam remnant splitting of momentum between a hadron and a quark/diquark jet, where \ttt{MSTP(94)} should now be used rather than \ttt{MSTP(92)}. \item In \ttt{PYDIFF} the recoiling gluon energy is calculated in a numerically more stable fashion. \end{Itemize} \item {\tsc{Pythia}} version 5.706, 25 August 1994: \begin{Itemize} \item New processes 167 and 168, $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \d^*$ and $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \d^*$, respectively, have been introduced. These contact interaction production processes of excited quarks complement the quark--gluon fusion ones in processes 147 and 148, and obey the same general rules, see section \ref{sss:qlstarclass}. \item The option \ttt{MSTP(57)=3} now also allows a dampening of $\pi^{\pm}$ parton distributions. \item A few minor errors have been corrected. \end{Itemize} \item {\tsc{Pythia}} version 5.707, 20 October 1994: \begin{Itemize} \item A major bug discovered in processes 121 and 122 (and thus also affecting 181, 182, 186 and 187), $\mathrm{g} \mathrm{g} (\mathrm{q} \overline{\mathrm{q}}) \to \mathrm{Q} \overline{\mathrm{Q}} \H$: the kinematics was incorrectly handed on to the Kunszt matrix elements. This affected the default option $\mathrm{Q} = \t$, but effects were especially dramatic when the alternative $\mathrm{Q} = \b$ was used. The choice of appropriate $Q^2$ scale for structure functions introduces a further uncertainty in cross sections for the processes above. So long as only $\t$ quarks are considered, the $\t$ mass is a reasonable choice, but for the $\mathrm{Q} = \b$ alternative this is presumably too low. Therefore new options have been introduced in \ttt{MSTP(39)}, with the default behaviour changed. \item Another important bug corrected in the calculation of the reduction of $\t\overline{\mathrm{t}}$ cross section when decay modes are forced. This occured when both $\t$ and $\overline{\mathrm{t}}$ produced a $\mathrm{W}$, and $\mathrm{W}^+$ and $\mathrm{W}^-$ decay modes were set differently. \end{Itemize} \item {\tsc{Pythia}} version 5.708, 25 October 1994: \begin{Itemize} \item A few further places changed to make processes 181, 182, 186 and 187 work (see version 5.707 above). \end{Itemize} \item {\tsc{Pythia}} version 5.709, 26 October 1994: \begin{Itemize} \item The matrix element for $\mathrm{f} \overline{\mathrm{f}} \to \mathrm{W}^+ \mathrm{W}^-$ has been replaced, using the formulae of \\ D. Bardin, M. Bilenky, D. Lehner, A. Olchevski and T. Riemann, CERN-TH.7295/94, \\ but with the dependence on the $\hat{t}$ variable not integrated out (D. Bardin, private communication). This avoids some problems encountered in the old expressions when one or both $\mathrm{W}$'s were far off the mass shell. \item Change in calls to \tsc{Pdflib}, so that the input $Q$ is always at least the $Q_{\mathrm{min}}$ of the respective set. \item Extra protection against infinite loops in PYSSPA. \end{Itemize} \item {\tsc{Pythia}} version 5.710, 27 January 1995: \begin{Itemize} \item The dimensions of the \ttt{HGZ} array in \ttt{PYRESD} has been expanded to avoid accidental writing outside the bounds. \item \ttt{VINT(41)-VINT(66)} are saved and restored in \ttt{PYSCAT}, for use in low-$p_{\perp}$ events, when beam remnant treatment has failed (with nonzero \ttt{MINT(57)}). \item The routine \ttt{PYSTGH} has been replaced by the routine \ttt{PYSTHG}. This contains an improved parametrization of the homogeneous evolution of an anomalous photon from some given initial scale. The argument \ttt{NF} of the \ttt{PYSTGH} routine has been removed; now $\Lambda$ is always automatically converted to the relevant $n_f$-flavour value from its 4-flavour one, at flavour thresholds. \end{Itemize} \item {\tsc{Pythia}} version 5.711, 20 February 1995: \begin{Itemize} \item New possibilities have been added to switch between electroweak couplings being expressed in terms of a running $\alpha_{\mathrm{em}}(Q^2)$ or in terms of a fixed Fermi constant $G_{\mrm{F}}$. This affects both decay widths and process cross sections, in the routines \ttt{PYINRE}, \ttt{PYRESD}, \ttt{PYWIDT} and \ttt{PYSIGH}. See \ttt{MSTP(8)} for details; default corresponds to old behaviour. \item The option \ttt{MSTP(37)=1}, with running quark masses in couplings to Higgs bosons, only works when $\alpha_{\mathrm{s}}$ is allowed to run (so one can define a $\Lambda$ value). Therefore a check has been introduced in \ttt{PYWIDT} and \ttt{PYSIGH} that the option \ttt{MSTP(37)=1} is only executed if additionally \ttt{MSTP(2)}$\geq 1$. \item Some non-physics changes have been made in the \ttt{RKBBV} and \ttt{STRUCTM} codes so as to avoid some (in principle harmless) compiler warnings. \end{Itemize} \item {\tsc{Pythia}} version 5.712, 15 March 1995: \begin{Itemize} \item A serious error has been corrected in the \ttt{MSTP(173)=1} option, i.e. when the program is run with user-defined weights that should compensate for a biased choice of variable beam energies. This both affected the relative admixture of low- and high-$p_{\perp}$ events and the total cross section obtained by Monte Carlo integration. (\ttt{PYRAND} changed.) \item In order to improve the flexibility and efficiency of the variable-energy option, the user should now set \ttt{PARP(174)} before the \ttt{PYINIT} call, and thereafter not change it. This allows \ttt{PARP(173)} weights of arbitrary size. (\ttt{PYRAND} and \ttt{PYMAXI} changed.) \item \ttt{MSTI(5)} (and \ttt{MINT(5)}) are now changed so they count the number of successfully generated events, rather than the number of tries made. This change only affects runs with variable energies, \ttt{MSTP(171)=1} and \ttt{MSTP(172)=2}, where \ttt{MSTI(61)=1} signals that a user-provided energy has been rejected in the weighting. This change also affects \ttt{PARI(2)}, which becomes the cross section per fully generated event. (\ttt{PYEVNT} changed.) \item The option \ttt{MSTP(14)=10} has now been extended so that it also works for deep inelastic scattering of an electron off a (real) photon, i.e. subprocess 10. What is obtained is a mixture of the photon acting as a vector meson and it acting as an anomalous state. This should therefore be the sum of what can be obtained with \ttt{MSTP(14)=2} and \ttt{=3}. It is distinct from \ttt{MSTP(14)=1} in that different sets are used for the parton distributions --- in \ttt{MSTP(14)=1} all the contributions to the photon distributions are lumped together, while they are split in VMD and anomalous parts for \ttt{MSTP(14)=10}. Also the beam remnant treatment is different, with a simple Gaussian distribution (at least by default) for \ttt{MSTP(14)=1} and the VMD part of \ttt{MSTP(14)=10}, but a powerlike distribution $\d k_{\perp}^2)/k_{\perp}^2$ between \ttt{PARP(15)} and $Q$ for the anomalous part of \ttt{MSTP(14)=10}. (\ttt{PYINIT}, \ttt{PYINPR} and \ttt{PYSTAT} changed.)\\ To access this option for $\mathrm{e}$ and $\gamma$ as incoming beams, it is only necessary to set \ttt{MSTP(14)=10} and keep \ttt{MSEL} at its default value. Unlike the corresponding option for $\gamma\mathrm{p}$ and $\gamma\gamma$, no cuts are overwritten, i.e. it is still the responsability of the user to set these appropriately. Those especially appropriate for DIS usage are \ttt{CKIN(21)-CKIN(22)} or \ttt{CKIN(23)-CKIN(24)} for the $x$ range (former or latter depending on which side is the incoming real photon), and \ttt{CKIN(35)-CKIN(36)} for the $Q^2$ range. A further new option has been added (in \ttt{PYKLIM}) to set the $W^2$ range as well, see \ttt{CKIN(39)-CKIN(40)}.\\ A warning about the usage of \tsc{Pdflib} for photons. So long as \ttt{MSTP(14)=1}, i.e. the photon is not split up, \tsc{Pdflib} is accessed by \ttt{MSTP(56)=2} and \ttt{MSTP(55)} the parton distribution set, as described in the manual. However, when the VMD and anomalous pieces are split, the VMD part is based on a rescaling of pion distributions by VMD factors (except for the SaS sets, that already come with a separate VMD piece). Therefore, to access \tsc{Pdflib} for \ttt{MSTP(14)=10}, it is not correct to set \ttt{MSTP(56)=2} and a photon distribution in \ttt{MSTP(55)}. Instead, one should put \ttt{MSTP(56)=2}, \ttt{MSTP(54)=2} and a pion distribution code in \ttt{MSTP(53)}, while \ttt{MSTP(55)} has no function. The anomalous part is still based on the SaS parametrization, with \ttt{PARP(15)} as main free parameter. \item A change has been made in \ttt{PYREMN} to reduce the possibility of infinite loops. \end{Itemize} \item {\tsc{Pythia}} version 5.713, 22 March 1995: \begin{Itemize} \item The SaS parton distributions of the photons are now available, see \cite{Sch95}. There are four new sets. These differ in that two use a $Q_0=0.6$~GeV and two a $Q_0=2$~GeV, and in that two use the DIS and two the $\overline{\mathrm{MS}}$ conventions for the dominant non-leading contributions. (However, the fits are formally still leading-order, in that not all next-to-leading contributions have been included.) New default is the SaS 1D set. Furthermore, for the definition of $F_2^{\gamma}$, additional terms appear that do not form part of the parton distributions itself. To partly take this into account, an additional doubling of the possibilities has been included. These eight possibilites can be accesed with \ttt{MSTP(55)}. The default value of \ttt{PARP(15)} has been changed from 0.5 to 0.6~GeV, for consistency with SaS 1D.\\ The generic routine \ttt{PYSTFU} has been rewritten to handle the interfacing. The old routines \ttt{PYSTAG}, \ttt{PYSTGS}, \ttt{PYDILN} and \ttt{PYSTHG} have been removed. Instead the routines of the \tsc{SaSgam} library have been inserted. In order to avoid any clashes, the routines \ttt{SAS***} have been renamed \ttt{PYG***}. Thus new routines are \ttt{PYGGAM}, \ttt{PYGVMD}, \ttt{PYGANO}, \ttt{PYGBEH} and \ttt{PYGDIR}. The common block \ttt{SASCOM} is renamed \ttt{PYINT8}. If you want to use the parton distributions for standalone purposes, you are encouraged to use the original \tsc{SaSgam} routines rather than going the way via the {\tsc{Pythia}} adaptations. \item \ttt{PYDOCU} has been corrected so that \ttt{PARI(2)} refers to the full cross section for $\gamma\mathrm{p}$ and $\gamma\gamma$ processes, rather than that of the latest subprocess considered. \item An additional check has been inserted into PYREMN. \end{Itemize} \item {\tsc{Pythia}} version 5.714, 22 March 1995: \begin{Itemize} \item Some minor modifications to \ttt{PYSTFU} and \ttt{PYGGAM} in the wake of the changes of the previous version. \end{Itemize} \item {\tsc{Pythia}} version 5.715, 24 April 1995: \begin{Itemize} \item An unfortunate choice of default values has been corrected: the old \ttt{MSTP(3)=2} value implied that $\Lambda_{\mrm{QCD}}$ was entirely based on the $\Lambda$ value of the proton structure function; also e.g. for $\e^+\e^-$ annihilation events. Thus the $\Lambda$ in \ttt{PARJ(81)} was overwritten, i.e. did not keep the value required by standard phenomenology, which typically gave too narrow jets. (While switching to \ttt{MSTP(3)=1} it worked fine.) In the modified option \ttt{MSTP(3)=2} this has been corrected, to better agree with user expectations. Since further changes were made in version 5.716, we refer below for additional comments. \item The form for \ttt{PTMANO}, the $p_{\perp\mathrm{min}}$ for anomalous processes, as used in \ttt{PYINPR} when processes are mixed for $\gamma\mathrm{p}$ or $\gamma\gamma$ events, has been updated to match (as well as can be expected) the SaS 1D photon distributions. \end{Itemize} \item {\tsc{Pythia}} version 5.716, 30 June 1995: \begin{Itemize} \item The strategy for the changes to $\Lambda$ in version 5.715 above have been modified for better transparency. Now \ttt{PARJ(81)} is used for resonance decays (including e.g. $\mathrm{Z}^0$ decay, from which it is determined), and \ttt{PARP(72)} for other timelike showers. \ttt{PARJ(81)} is not overwritten \ttt{for MSTP(3)=2}, but only for \ttt{=3}. Changes affect \ttt{PYINIT}, \ttt{PYEVNT} and \ttt{PYRESD}. \item A new multiplicative factor has been introduced for the $Q^2$ scale choice of the hard scattering in \ttt{PYSIGH}, affecting parton distributions and $\alpha_{\mathrm{s}}$, see \ttt{PARP(34)}. \item \ttt{PYREMN} has been corrected for occasional too large boost factors. \item An error in \ttt{PYSIGH} for process 148 has been corrected. \item The \ttt{MSTP(62)=1} option of \ttt{PYSSPA} is modified to avoid division by zero. \item Header has been updated with WWW-information. \end{Itemize} \item {\tsc{Pythia}} version 5.717, 23 August 1995: \begin{Itemize} \item \ttt{MIN1}, \ttt{MIN2}, \ttt{MAX1}, \ttt{MAX2}, \ttt{MINA} and \ttt{MAXA} in \ttt{PYSIGH} have had an extra \ttt{M} prefixed to avoid confusion with Fortran functions. \item Protect against \ttt{MDCY(0,1)} being accessed in \ttt{PYSIGH}. \item Protect against \ttt{THB=0} in \ttt{PYRAND}. \item Protect against \ttt{YSTMAX-YSTMIN = 0} in \ttt{PYSIGH}. \item Check for moved leptoquark at beginning of \ttt{PYRESD} just like for other particles with colour. \end{Itemize} \end{Enumerate} \subsection{Program Installation} \label{ss:install} Several `authorized' sources of the programs exist. The `master copy' of the programs is the one found on my World Wide Web homepage\\ \ttt{http://thep.lu.se/tf2/staff/torbjorn/Welcome.html}\\ There you have: \begin{tabbing} ~~~ \= \ttt{jetset74.f}~~~~~~~ \= the {\tsc{Jetset}} code, \\ \> \ttt{pythia57.f} \> the {\tsc{Pythia}} code, \\ \> \ttt{pythia57.tex} \> this common {\tsc{Pythia/Jetset}} manual, and \\ \> \ttt{update57.notes} \> plain text update notes to the manual. \end{tabbing} In addition to these, one may also find older versions of the program and manuals, sample main programs and other pieces of related software, and other physics papers. The lack of stable versions may make it less convenient to rely on the above files. New versions are introduced in the general distribution of the CERN program library, maybe once a year. These versions are better checked before release, and should be useful for most applications. However, clearly, they may be less up-to-date. Read the CERN Computer Newsletter for announcements. Copies of the programs are also available via anonymous ftp, e.g. from the asisftp server at CERN. The programs are written entirely in standard Fortran 77, and should run on any machine with such a compiler. To a first approximation, program compilation should therefore be straightforward. Unfortunately, experience with many different compilers has been uniform: the options available for obtaining optimized code actually produce erroneous code (e.g. operations inside \ttt{DO} loops are moved out before them, where some of the variables have not yet been properly set). Therefore the general advice is to use a low optimization level. Note that this is often not the default setting. \ttt{SAVE} statements have been included in accordance with the Fortran standard. Since most ordinary machines take \ttt{SAVE} for granted, this part is not particularly well tried out, however. All default settings and particle and process data are stored in \ttt{BLOCK DATA LUDATA} for {\tsc{Jetset}} and \ttt{BLOCK DATA PYDATA} for {\tsc{Pythia}}. These subprograms must be linked for a proper functioning of the other routines. On some machines this is not done automatically but must be forced by you, in particular if {\tsc{Jetset}} and {\tsc{Pythia}} are maintained as libraries from which routines are to be loaded only when they are needed. In this connection we note that the library approach does not give any significant space advantages over a loading of the packages as a whole, since a normal run will call on most of the routines anyway, directly or indirectly. Since most machines in current use are 32-bit ones, this is the precision normally assumed. A few pieces of code have therefore had to be written in double precision. As a rule of thumb, double-precision variables have as first character \ttt{D}, but there are a few exceptions. For applications at very high energies, such as LHC, the use of single precision for any real variable is a problem. It might then be necessary to rewrite the program completely, i.e. to have a declaration \ttt{IMPLICIT DOUBLE PRECISION(A-H,O-Z)} at the beginning of each subprogram, and to change all real constants to double precision. Needless to say, the latter is a major undertaking. In some cases, shortcuts are available. On the IBM, for instance, the \ttt{AUTODBL} compiler option for automatic precision doubling works fine, provided only that an even number of integers precede real numbers in common blocks. In {\tsc{Jetset}} you therefore need to introduce an additional integer variable (\ttt{NPAD}, say) directly after \ttt{N} in the \ttt{LUJETS} common block, and in {\tsc{Pythia}} an additional integer (\ttt{MSEPAD}) after \ttt{MSEL} in the \ttt{PYSUBS} common block. Some pieces of code will then actually run in quadruple precision. A test program, \ttt{LUTEST}\label{p:LUTEST}, is included in the {\tsc{Jetset}} package. It is disguised as a subroutine, so you have to run a main program \begin{verbatim} CALL LUTEST(1) END \end{verbatim} This program will generate six hundred events of different types, under a variety of conditions. If {\tsc{Jetset}} has not been properly installed, this program is likely to crash, or at least generate a number of erroneous events. This will then clearly be marked in the output, which otherwise will just contain a few sample event listings and a table of the number of different particles produced. To switch off the output of normal events and final table, use \ttt{LUTEST(0)} instead of \ttt{LUTEST(1)}. The final tally of errors detected should read 0. In exactly the same vein, a test program \ttt{PYTEST}\label{p:PYTEST} comes with the {\tsc{Pythia}} package. You then have to run a program \begin{verbatim} CALL PYTEST(1) END \end{verbatim} As before the alternative \ttt{PYTEST(0)} will give a less extensive listing. No errors should appear during execution. \subsection{Program Philosophy} The Monte Carlo programs are built as slave systems, i.e. you, the user, have to supply the main program. From this the various subroutines are called on to execute specific tasks, after which control is returned to the main program. Some of these tasks may be very trivial, whereas the `high-level' routines by themselves may make a large number of subroutine calls. Many routines are not intended to be called directly by you, but only from higher-level routines such as \ttt{LUEXEC}, \ttt{LUEEVT}, \ttt{PYINIT} or \ttt{PYEVNT}. Basically, this means that there are three ways by which you communicate with the programs. First, by setting common block variables, you specify the details of how the programs should perform specific tasks, i.e. which subprocesses should be generated (for {\tsc{Pythia}}), which particle masses should be assumed, which coupling constants used, which fragmentation scenarios, and so on with hundreds of options and parameters. Second, by calling subroutines you tell the programs to generate events according to the rules established above. Normally there are few subroutine arguments, and those are usually related to details of the physical situation, such as what c.m. energy to assume for events. Third, you can either look at the common block \ttt{LUJETS} to extract information on the generated event, or you can call on various functions and subroutines to analyse the event further for you. It should be noted that, while the physics content is obviously at the centre of attention, the {\tsc{Jetset/Pythia}} package also contains a very extensive setup of auxiliary service routines. The hope is that this will provide a comfortable working environment, where not only events are generated, but where you also linger on to perform a lot of the subsequent studies. Of course, for detailed studies, it may be necessary to interface the output directly to a detector simulation program. The general rule is that all routines have names that are six characters long, beginning with \ttt{LU} for {\tsc{Jetset}} routines and \ttt{PY} for {\tsc{Pythia}} ones. Real-valued functions in {\tsc{Jetset}} begin with \ttt{UL} instead. There are three exceptions to both the length and the initial character rules: \ttt{KLU}, \ttt{PLU} and \ttt{RLU}. The former two functions are strongly coupled to the \ttt{K} and \ttt{P} matrices in the \ttt{LUJETS} common block, the latter uses \ttt{R} to emphasize the r\^ole as a random-number generator. Also common block names are six characters long and start with \ttt{LU} or \ttt{PY}. On the issue of initialization, {\tsc{Jetset}} and {\tsc{Pythia}} behave quite differently. Most {\tsc{Jetset}} routines work without any initialization (except for the one implied by the presence of \ttt{BLOCK DATA LUDATA}, see above), i.e. each event and each task stand on their own. Current common block values are used to perform the tasks in specific ways, and those rules can be changed from one event to the next (or even within the generation of one and the same event) without any penalty. The random-number generator is initialized at the first call, but usually this is transparent. Therefore the two {\tsc{Jetset}} routines \ttt{LUEEVT} (and some of the routines called by it) and \ttt{LUONIA} are basically the only ones to contain some elements of initialization, where there are a few advantages if events are generated in a coherent fashion, but even here the penalty for not doing it is small. In {\tsc{Pythia}}, on the other hand, a sizeable amount of initialization is performed in the \ttt{PYINIT} call, and thereafter the events generated by \ttt{PYEVNT} all obey the rules established at that point. Therefore common block variables that specify methods to be used have to be set before the \ttt{PYINIT} call and then not be changed afterwards, with few exceptions. Of course, it is possible to perform several \ttt{PYINIT} calls in the same run, but there is a significant time overhead involved, so this is not something one would do for each new event. Apart from writing a title page, giving a brief initialization information, printing error messages if need be, and responding to explicit requests for listings, all tasks of the programs are performed `silently'. All output is directed to unit \ttt{MSTU(11)}, by default 6, and it is up to you to set this unit open for write. The only exceptions are \ttt{RLUGET}, \ttt{RLUSET} and \ttt{LUUPDA} where, for obvious reasons, the input/output file number is specified at each call. Here you again have to see to it that proper read/write access is set. The programs are extremely versatile, but the price to be paid for this is having a large number of adjustable parameters and switches for alternative modes of operation. No single user is ever likely to need more than a fraction of the available options. Since all these parameters and switches are assigned sensible default values, there is no reason to worry about them until the need arises. Unless explicitly stated (or obvious from the context) all switches and parameters can be changed independently of each other. One should note, however, that if only a few switches/parameters are changed, this may result in an artificially bad agreement with data. Many disagreements can often be cured by a subsequent retuning of some other parameters of the model, in particular those that were once determined by a comparison with data in the context of the default scenario. For example, for $\e^+\e^-$ annihilation, such a retuning could involve one QCD parameter ($\alpha_{\mathrm{s}}$ or $\Lambda$), the longitudinal fragmentation function, and the average transverse fragmentation momentum. The programs contain a number of checks that requested processes have been implemented, that flavours specified for jet systems make sense, that the energy is sufficient to allow hadronization, that the memory space in \ttt{LUJETS} is large enough, etc. If anything goes wrong that the program can catch (obviously this may not always be possible), an error message will be printed and the treatment of the corresponding event will be cut short. In serious cases, the program will abort. As long as no error messages appear on the output, it may not be worthwhile to look into the rules for error checking, but if but one message appears, it should be enough cause for alarm to receive prompt attention. Also warnings are sometimes printed. These are less serious, and the experienced user might deliberately do operations which go against the rules, but still can be made to make sense in their context. Only the first few warnings will be printed, thereafter the program will be quiet. By default, the program is set to stop execution after ten errors, after printing the last erroneous event. It must be emphasized that not all errors will be caught. In particular, one tricky question is what happens if an integer-valued common block switch or subroutine/function argument is used with a value that is not defined. In some subroutine calls, a prompt return will be expedited, but in most instances the subsequent action is entirely unpredictable, and often completely haywire. The same goes for real-valued variables that are assigned values outside the physically sensible range. One example will suffice here: if \ttt{PARJ(2)} is defined as the $\mathrm{s} / \u$ suppression factor, a value $>1$ will not give more profuse production of $\mathrm{s}$ than of $\u$, but actually a spillover into $\c$ production. Users, beware! \subsection{Manual Conventions} In the manual parts of this report, some conventions are used. All names of subprograms, common blocks and variables are given in upper-case `typewriter' style, e.g. \ttt{MSTP(111)=0}. Also program examples are given in this style. If a common block variable must have a value set at the beginning of execution, then a default value is stored in one of the block data subprograms \ttt{LUDATA} and \ttt{PYDATA}. Such a default value is usually indicated by a `(D=\ldots)' immediately after the variable name, e.g. \begin{entry} \iteme{MSTJ(1) :} (D=1) choice of fragmentation scheme. \end{entry} All variables in the {\tsc{Jetset}} common blocks (with very few exceptions, clearly marked) can be freely changed from one event to the next, or even within the treatment of one single event. In the {\tsc{Pythia}} common blocks the situation is more complicated. The values of many switches and parameters are used already in the \ttt{PYINIT} call, and cannot be changed after that. The problem is mentioned in the preamble to the afflicted common blocks, which in particular means \ttt{/PYPARS/} and \ttt{/PYSUBS/}. For the variables which may still be changed from one event to the next, a `(C)' is added after the `(D=\ldots)' statement. Normally, variables internal to the program are kept in separate common blocks and arrays, but in a few cases such internal variables appear among arrays of switches and parameters, mainly for historical reasons. These are denoted by `(R)' for variables you may want to read, because they contain potentially interesting information, and by `(I)' for purely internal variables. In neither case may the variables be changed by you. In the description of a switch, the alternatives that this switch may take are often enumerated, e.g. \begin{entry} \iteme{MSTJ(1) :} (D=1) choice of fragmentation scheme. \begin{subentry} \iteme{= 0 :} no jet fragmentation at all. \iteme{= 1 :} string fragmentation according to the Lund model. \iteme{= 2 :} independent fragmentation, according to specification in \ttt{MSTJ(2)} and \ttt{MSTJ(3)}. \end{subentry} \end{entry} If you then use any value other than 0, 1 or 2, results are unpredictable. The action could even be different in different parts of the program, depending on the order in which the alternatives are identified. It is also up to you to choose physically sensible values for parameters: there is no check on the allowed ranges of variables. We gave an example of this at the end of the preceding section. Subroutines you are expected to use are enclosed in a box at the point where they are defined: \drawbox{CALL LULIST(MLIST)} \boxsep This is followed by a description of input or output parameters. The difference between input and output is not explicitly marked, but should be obvious from the context. In fact, the event-analysis routines of section \ref{ss:evanrout} return values, while all the rest only have input variables. Routines that are only used internally are not boxed in. However, we use boxes for all common blocks, so as to enhance the readability. \subsection{Getting Started with JETSET} \label{ss:JETstarted} As a first example, assume that you want to study the production of $\u \overline{\mathrm{u}}$ 2-jet systems at 20 GeV energy. To do this, write a main program \begin{verbatim} CALL LU2ENT(0,2,-2,20.) CALL LULIST(1) END \end{verbatim} and run this program, linked together with {\tsc{Jetset}}. The routine \ttt{LU2ENT} is specifically intended for storing two entries (jets or particles). The first argument (0) is a command to perform fragmentation and decay directly after the entries have been stored, the second and third that the two entries are $\u$ (2) and $\overline{\mathrm{u}}$ ($-2$), and the last that the c.m. energy of the pair is 20 GeV. When this is run, the resulting event is stored in the \ttt{LUJETS} common block. This information can then be read out by you. No output is produced by \ttt{LU2ENT} itself, except for a title page which appears once for every {\tsc{Jetset/Pythia}} run. Instead the second command, to \ttt{LULIST}, provides a simple visible summary of the information stored in \ttt{LUJETS}. The argument (1) indicates that the short version should be used, which is suitable for viewing the listing directly on an 80-column terminal screen. It might look as shown here. \begin{verbatim} Event listing (summary) I particle/jet KS KF orig p_x p_y p_z E m 1 (u) A 12 2 0 0.000 0.000 10.000 10.000 0.006 2 (u~) V 11 -2 0 0.000 0.000 -10.000 10.000 0.006 3 (string) 11 92 1 0.000 0.000 0.000 20.000 20.000 4 (rho+) 11 213 3 0.098 -0.154 2.710 2.856 0.885 5 (rho-) 11 -213 3 -0.227 0.145 6.538 6.590 0.781 6 pi+ 1 211 3 0.125 -0.266 0.097 0.339 0.140 7 (Sigma0) 11 3212 3 -0.254 0.034 -1.397 1.855 1.193 8 (K*+) 11 323 3 -0.124 0.709 -2.753 2.968 0.846 9 p~- 1 -2212 3 0.395 -0.614 -3.806 3.988 0.938 10 pi- 1 -211 3 -0.013 0.146 -1.389 1.403 0.140 11 pi+ 1 211 4 0.109 -0.456 2.164 2.218 0.140 12 (pi0) 11 111 4 -0.011 0.301 0.546 0.638 0.135 13 pi- 1 -211 5 0.089 0.343 2.089 2.124 0.140 14 (pi0) 11 111 5 -0.316 -0.197 4.449 4.467 0.135 15 (Lambda0) 11 3122 7 -0.208 0.014 -1.403 1.804 1.116 16 gamma 1 22 7 -0.046 0.020 0.006 0.050 0.000 17 K+ 1 321 8 -0.084 0.299 -2.139 2.217 0.494 18 (pi0) 11 111 8 -0.040 0.410 -0.614 0.751 0.135 19 gamma 1 22 12 0.059 0.146 0.224 0.274 0.000 20 gamma 1 22 12 -0.070 0.155 0.322 0.364 0.000 21 gamma 1 22 14 -0.322 -0.162 4.027 4.043 0.000 22 gamma 1 22 14 0.006 -0.035 0.422 0.423 0.000 23 p+ 1 2212 15 -0.178 0.033 -1.343 1.649 0.938 24 pi- 1 -211 15 -0.030 -0.018 -0.059 0.156 0.140 25 gamma 1 22 18 -0.006 0.384 -0.585 0.699 0.000 26 gamma 1 22 18 -0.034 0.026 -0.029 0.052 0.000 sum: 0.00 0.000 0.000 0.000 20.000 20.000 \end{verbatim} (A few blanks have been removed between the columns to make it fit into the format of this text.) Look in the particle/jet column and note that the first two lines are the original $\u$ and $\overline{\mathrm{u}}$, where `bar' is actually written `$\sim$' to save space in longer names. The parentheses enclosing the names, `\ttt{(u)}' and `\ttt{(u\~{})}', are there as a reminder that these jets actually have been allowed to fragment. The jets are still retained so that event histories can be studied. Also note that the \ttt{KF} (flavour code) column contains 2 in the first line and $-2$ in the second. These are the codes actually stored to denote the presence of a $\u$ and a $\overline{\mathrm{u}}$, cf. the \ttt{LU2ENT} call, while the names written are just conveniences used when producing visible output. The \ttt{A} and \ttt{V} near the end of the particle/jet column indicate the beginning and end of a string (or cluster, or independent fragmentation) parton system; any intermediate entries belonging to the same system would have had an \ttt{I} in that column. (This gives a poor man's representation of an up-down arrow, $\updownarrow$.) In the \ttt{orig} (origin) column, the zeros indicate that $\u$ and $\overline{\mathrm{u}}$ are two initial entries. The subsequent line, number 3, denotes the fragmenting $\u \overline{\mathrm{u}}$ string system as a whole, and has origin 1, since the first parton of this string system is entry number 1. The particles in lines 4--10 have origin 3 to denote that they come directly from the fragmentation of this string. In string fragmentation it is not meaningful to say that a particle comes from only the $\u$ quark or only the $\overline{\mathrm{u}}$ one. It is the string system as a whole that gives a $\rho^+$, a $\rho^-$, a $\pi^+$, a $\Sigma^0$, a $\mathrm{K}^{*+}$, a $\overline{\mathrm{p}}^-$, and a $\pi^-$. Note that some of the particle names are again enclosed in parentheses, indicating that these particles are not present in the final state either, but have decayed further. Thus the $\pi^-$ in line 13 and the $\pi^0$ in line 14 have origin 5, as an indication that they come from the decay of the $\rho^-$ in line 5. Only the names not enclosed in parentheses remain at the end of the fragmentation/decay chain, and are thus experimentally observable. The actual status code used to distinguish between different classes of entries is given in the \ttt{KS} column; codes in the range 1--10 correspond to remaining entries, and those above 10 to those that have fragmented or decayed. The columns with \ttt{p\_x}, \ttt{p\_y}, \ttt{p\_z}, \ttt{E} and \ttt{m} are quite self-explanatory. All momenta, energies and masses are given in units of GeV, since the speed of light is taken to be $c = 1$. Note that energy and momentum are conserved at each step of the fragmentation/decay process (although there exist options where this is not true). Also note that the $z$ axis plays the r\^ole of preferred direction, along which the original partons are placed. The final line is intended as a quick check that nothing funny happened. It contains the summed charge, summed momentum, summed energy and invariant mass of the final entries at the end of the fragmentation/decay chain, and the values should agree with the input implied by the \ttt{LU2ENT} arguments. (In fact, warnings would normally appear on the output if anything untoward happened, but that is another story.) The above example has illustrated roughly what information is to be had in the event record, but not so much about how it is stored. This is better seen by using a 132-column format for listing events. Try e.g. the following program \begin{verbatim} CALL LU3ENT(0,1,21,-1,30.,0.9,0.7) CALL LULIST(2) CALL LUEDIT(3) CALL LULIST(2) END \end{verbatim} where a 3-jet $\d \mathrm{g} \overline{\mathrm{d}}$ event is generated in the first line and listed in the second. This listing will contain the numbers as directly stored in the common block \ttt{LUJETS} \begin{verbatim} COMMON/LUJETS/N,K(4000,5),P(4000,5),V(4000,5) \end{verbatim} For particle \ttt{I}, \ttt{K(I,1)} thus gives information on whether or not a jet or particle has fragmented or decayed, \ttt{K(I,2)} gives the particle code, \ttt{K(I,3)} its origin, \ttt{K(I,4)} and \ttt{K(I,5)} the position of fragmentation/decay products, and \ttt{P(I,1})--\ttt{P(I,5)} momentum, energy and mass. The number of lines in current use is given by \ttt{N}, i.e. 1 $\leq$ \ttt{I} $\leq$ \ttt{N}. The \ttt{V} matrix contains decay vertices; to view those \ttt{LULIST(3)} has to be used. It is important to learn the rules for how information is stored in \ttt{LUJETS}. The third line in the program illustrates another important point about {\tsc{Jetset}}: a number of routines are available for manipulating the event record after the event has been generated. Thus \ttt{LUEDIT(3)} will remove everything except stable charged particles, as shown by the result of the second \ttt{LULIST} call. More advanced possibilities include things like sphericity or clustering routines. Apart from the input arguments of subroutine calls, control on the doings of {\tsc{Jetset}} may be imposed via the \ttt{LUDAT1}, \ttt{LUDAT2}, \ttt{LUDAT3} and \ttt{LUDAT4} common blocks. Here sensible default values are always provided. A user might want to switch off all particle decays by putting \ttt{MSTJ(21)=0} or increase the $\mathrm{s} / \u$ ratio in fragmentation by putting \ttt{PARJ(2)=0.40}, to give but two examples. It is by exploring the possibilities offered here that {\tsc{Jetset}} can be turned into an extremely versatile tool, even if all the nice physics is already present in the default values. As a final, semirealistic example, assume that the $p_{\perp}$ spectrum of $\pi^+$ particles is to be studied in 91.2 GeV $\e^+\e^-$ annihilation events, where $p_{\perp}$ is to be defined with respect to the sphericity axis. Using the HBOOK package (version 4, watch out for version- or installation-specific differences) for histogramming, a complete program might look like \begin{verbatim} C...Common blocks. COMMON/LUJETS/N,K(4000,5),P(4000,5),V(4000,5) COMMON/PAWC/HMEMOR(10000) C...Reserve histogram memory and book histograms. CALL HLIMIT(10000) CALL HBOOK1(1,'pT spectrum of pi+',100,0.,5.,0.) C...Number of events to generate. Loop over events. NEVT=100 DO 110 IEVT=1,NEVT C...Generate event. List first one. CALL LUEEVT(0,91.2) IF(IEVT.EQ.1) CALL LULIST(1) C...Find sphericity axis and rotate event so sphericity along z axis. CALL LUSPHE(SPH,APL) CALL LUEDIT(31) C...Loop over all particles, but skip if not pi+. DO 100 I=1,N IF(K(I,2).NE.211) GOTO 100 C...Calculate pT and fill in histogram. PT=SQRT(P(I,1)**2+P(I,2)**2) CALL HF1(1,PT,1.) C...End of particle and event loops. 100 CONTINUE 110 CONTINUE C...Normalize histogram properly and list it. CALL HOPERA(1,'+',1,1,20./NEVT,0.) CALL HISTDO END \end{verbatim} Study this program, try to understand what happens at each step, and run it to check that it works. You should then be ready to look at the relevant sections of this report and start writing your own programs. \subsection{Getting Started with PYTHIA} \label{ss:PYTstarted} A {\tsc{Pythia}} run has to be more strictly organized than a {\tsc{Jetset}} one, in that it is necessary to initialize the generation before events can be generated, and in that it is not possible to change switches and parameters freely during the course of the run. A fairly precise recipe for how a run should be structured can therefore be given. Thus, the usage of {\tsc{Pythia}} can be subdivided into three steps. \begin{Enumerate} \item The initialization step. It is here that all the basic characteristics of the coming generation are specified. The material in this section includes the following. \begin{Itemize} \item Common blocks, at least the following, and maybe some more: \begin{verbatim} COMMON/LUJETS/N,K(4000,5),P(4000,5),V(4000,5) COMMON/LUDAT1/MSTU(200),PARU(200),MSTJ(200),PARJ(200) COMMON/PYSUBS/MSEL,MSUB(200),KFIN(2,-40:40),CKIN(200) COMMON/PYPARS/MSTP(200),PARP(200),MSTI(200),PARI(200) \end{verbatim} \vspace{-\baselineskip} \item Selection of required processes. Some fixed `menus' of subprocesses can be selected with different \ttt{MSEL} values, but with {\tt MSEL}=0 it is possible to compose `\`a la carte', using the subprocess numbers. To generate processes 14, 18 and 29, for instance, one needs \begin{verbatim} MSEL=0 MSUB(14)=1 MSUB(18)=1 MSUB(29)=1 \end{verbatim} \vspace{-\baselineskip} \item Selection of kinematics cuts in the \ttt{CKIN} array. To generate hard scatterings with 5~GeV $\leq p_{\perp} \leq$ 10~GeV, for instance, use \begin{verbatim} CKIN(3)=5. CKIN(4)=10. \end{verbatim} \vspace{-\baselineskip} Unfortunately, initial- and final-state radiation will shift around the kinematics of the hard scattering, making the effects of cuts less predictable. One therefore always has to be very careful that no desired event configurations are cut out. \item Definition of underlying physics scenario, e.g. top mass. \item Selection of parton-distribution sets, $Q^2$ definitions, and all other details of the generation. \item Switching off of generator parts not needed for toy simulations, e.g. fragmentation for parton level studies. \item Initialization of the event generation procedure. Here kinematics is set up, maxima of differential cross sections are found for future Monte Carlo generation, and a number of other preparatory tasks carried out. Initialization is performed by \ttt{PYINIT}, which should be called only after the switches and parameters above have been set to their desired values. The frame, the beam particles and the energy have to be specified. \begin{verbatim} CALL PYINIT('CMS','p','pbar',1800.) \end{verbatim} \vspace{-\baselineskip} \item Any other initial material required by the user, e.g. histogram booking. \end{Itemize} \item The generation loop. It is here that events are generated and studied. It includes the following tasks: \begin{Itemize} \item Generation of the next event, with \begin{verbatim} CALL PYEVNT \end{verbatim} \vspace{-\baselineskip} \item Printing of a few events, to check that everything is working as planned, with \begin{verbatim} CALL LULIST(1) \end{verbatim} \vspace{-\baselineskip} \item An analysis of the event for properties of interest, either directly reading out information from the \ttt{LUJETS} common block or making use of a number of utility routines in {\tsc{Jetset}}. \item Saving of events on tape, or interfacing to detector simulation. \end{Itemize} \item The finishing step. Here the tasks are: \begin{Itemize} \item Printing a table of deduced cross sections, obtained as a by-product of the Monte Carlo generation activity, with the command \begin{verbatim} CALL PYSTAT(1) \end{verbatim} \vspace{-\baselineskip} \item Printing histograms and other user output. \end{Itemize} \end{Enumerate} To illustrate this structure, imagine a toy example, where one wants to simulate the production of a 300 GeV Higgs particle. In {\tsc{Pythia}}, a program for this might look something like the following. \begin{verbatim} C...Common blocks. COMMON/LUJETS/N,K(4000,5),P(4000,5),V(4000,5) COMMON/LUDAT1/MSTU(200),PARU(200),MSTJ(200),PARJ(200) COMMON/LUDAT2/KCHG(500,3),PMAS(500,4),PARF(2000),VCKM(4,4) COMMON/LUDAT3/MDCY(500,3),MDME(2000,2),BRAT(2000),KFDP(2000,5) COMMON/PYSUBS/MSEL,MSUB(200),KFIN(2,-40:40),CKIN(200) COMMON/PYPARS/MSTP(200),PARP(200),MSTI(200),PARI(200) COMMON/PAWC/HBOOK(10000) C...Number of events to generate. Switch on proper processes. NEV=1000 MSEL=0 MSUB(102)=1 MSUB(123)=1 MSUB(124)=1 C...Select t and H masses and kinematics cuts in mass. PMAS(6,1)=140. PMAS(25,1)=300. CKIN(1)=290. CKIN(2)=310. C...For simulation of hard process only: cut out unnecessary tasks. MSTP(61)=0 MSTP(71)=0 MSTP(81)=0 MSTP(111)=0 C...Initialize and list partial widths. CALL PYINIT('CMS','p','p',16000.) CALL PYSTAT(2) C...Book histograms. CALL HLIMIT(10000) CALL HBOOK1(1,'Higgs mass',50,275.,325.,0.) C...Generate events. Look at first few. DO 200 IEV=1,NEV CALL PYEVNT IF(IEV.LE.3) CALL LULIST(1) C...Loop over particles to find Higgs and histogram its mass. DO 100 I=1,N 100 IF(K(I,2).EQ.25) HMASS=P(I,5) CALL HF1(1,HMASS,1.) 200 CONTINUE C...Print cross sections and histograms. CALL PYSTAT(1) CALL HISTDO END \end{verbatim} Here 102, 123 and 124 are the three main Higgs production graphs $\mathrm{g} \mathrm{g} \rightarrow \H$, $\mathrm{Z} \mathrm{Z} \rightarrow \H$, and $\mathrm{W} \mathrm{W} \rightarrow \H$, and \ttt{MSUB(ISUB)=1} is the command to switch on process \ttt{ISUB}. Full freedom to combine subprocesses `\`a la carte' is ensured by \ttt{MSEL=0}; ready-made `menus' can be ordered with other \ttt{MSEL} numbers. The \ttt{PMAS} commands set the masses of the top quark and the Higgs itself, and the \ttt{CKIN} variables the desired mass range of the Higgs --- a Higgs with a 300 GeV nominal mass actually has a fairly broad Breit--Wigner type mass distribution. The \ttt{MSTP} switches that come next are there to modify the generation procedure, in this case to switch off initial- and final-state radiation, multiple interactions among beam jets, and fragmentation, to give only the `parton skeleton' of the hard process. The \ttt{PYINIT} call initializes {\tsc{Pythia}}, by finding maxima of cross sections, recalculating the Higgs decay properties (which depend on the Higgs mass), etc. The decay properties can be listed with \ttt{PYSTAT(2)}. Inside the event loop, \ttt{PYEVNT} is called to generate an event, and \ttt{LULIST(1)} to list the event. The information used by \ttt{LULIST(1)} is the event record, stored in the common block \ttt{LUJETS}. Here one finds all produced particles, both final and intermediate ones, with information on particle species and event history (\ttt{K} array), particle momenta (\ttt{P} array) and production vertices (\ttt{V} array). In the loop over all particles produced, \ttt{1} through \ttt{N}, the Higgs particle is found by its code, \ttt{K(I,2)=25}, and its mass is stored in \ttt{P(I,5)}. After all events have been generated, \ttt{PYSTAT(1)} gives a summary of the number of events generated in the various allowed channels, and the inferred cross sections. In the run above, a typical event listing might look like the following. \begin{verbatim} Event listing (summary) I particle/jet KF p_x p_y p_z E m 1 !p+! 2212 0.000 0.000 8000.000 8000.000 0.938 2 !p+! 2212 0.000 0.000-8000.000 8000.000 0.938 ====================================================================== 3 !g! 21 -0.505 -0.229 28.553 28.558 0.000 4 !g! 21 0.224 0.041 -788.073 788.073 0.000 5 !g! 21 -0.505 -0.229 28.553 28.558 0.000 6 !g! 21 0.224 0.041 -788.073 788.073 0.000 7 !H0! 25 -0.281 -0.188 -759.520 816.631 300.027 8 !W+! 24 120.648 35.239 -397.843 424.829 80.023 9 !W-! -24 -120.929 -35.426 -361.677 391.801 82.579 10 !e+! -11 12.922 -4.760 -160.940 161.528 0.001 11 !nu_e! 12 107.726 39.999 -236.903 263.302 0.000 12 !s! 3 -62.423 7.195 -256.713 264.292 0.199 13 !c~! -4 -58.506 -42.621 -104.963 127.509 1.350 ====================================================================== 14 (H0) 25 -0.281 -0.188 -759.520 816.631 300.027 15 (W+) 24 120.648 35.239 -397.843 424.829 80.023 16 (W-) -24 -120.929 -35.426 -361.677 391.801 82.579 17 e+ -11 12.922 -4.760 -160.940 161.528 0.001 18 nu_e 12 107.726 39.999 -236.903 263.302 0.000 19 s A 3 -62.423 7.195 -256.713 264.292 0.199 20 c~ V -4 -58.506 -42.621 -104.963 127.509 1.350 21 ud_1 A 2103 -0.101 0.176 7971.328 7971.328 0.771 22 d V 1 -0.316 0.001 -87.390 87.390 0.010 23 u A 2 0.606 0.052 -0.751 0.967 0.006 24 uu_1 V 2203 0.092 -0.042-7123.668 7123.668 0.771 ====================================================================== sum: 2.00 0.00 0.00 0.00 15999.98 15999.98 \end{verbatim} The above event listing is abnormally short, in part because some columns of information were removed to make it fit into this text, in part because all initial- and final-state QCD radiation, all non-trivial beam jet structure, and all fragmentation was inhibited in the generation. Therefore only the skeleton of the process is visible. In lines 1 and 2 one recognizes the two incoming protons. In lines 3 and 4 are incoming partons before initial-state radiation and in 5 and 6 after --- since there is no such radiation they coincide here. Line 7 shows the Higgs produced by $\mathrm{g} \mathrm{g}$ fusion, 8 and 9 its decay products and 10--13 the second-step decay products. Up to this point lines give a summary of the event history, indicated by the exclamation marks that surround particle names (and also reflected in the \ttt{K(I,1)} code, not shown). From line 14 onwards come the particles actually produced in the final states, first in lines 14--16 particles that subsequently decayed, which have their names surrounded by brackets, and finally the particles and jets left in the end, including beam remnants. Here this also includes a number of unfragmented jets, since fragmentation was inhibited. Ordinarily, the listing would have gone on for a few hundred more lines, with the particles produced in the fragmentation and their decay products. The final line gives total charge and momentum, as a convenient check that nothing unexpected happened. The first column of the listing is just a counter, the second gives the particle name and information on status and string drawing (the \ttt{A} and \ttt{V}), the third the particle-flavour code (which is used to give the name), and the subsequent columns give the momentum components. One of the main problems is to select kinematics efficiently. Imagine for instance that one is interested in the production of a single $\mathrm{Z}$ with a transverse momentum in excess of 50 GeV. If one tries to generate the inclusive sample of $\mathrm{Z}$ events, by the basic production graphs $\mathrm{q} \overline{\mathrm{q}} \rightarrow \mathrm{Z}$, then most events will have low transverse momenta and will have to be discarded. That any of the desired events are produced at all is due to the initial-state generation machinery, which can build up transverse momenta for the incoming $\mathrm{q}$ and $\overline{\mathrm{q}}$. However, the amount of initial-state radiation cannot be constrained beforehand. To increase the efficiency, one may therefore turn to the higher-order processes $\mathrm{q} \mathrm{g} \rightarrow \mathrm{Z} \mathrm{q}$ and $\mathrm{q} \overline{\mathrm{q}} \rightarrow \mathrm{Z} \mathrm{g}$, where already the hard subprocess gives a transverse momentum to the $\mathrm{Z}$. This transverse momentum can be constrained as one wishes, but again initial- and final-state radiation will smear the picture. If one were to set a $p_{\perp}$ cut at 50 GeV for the hard-process generation, those events where the $\mathrm{Z}$ was given only 40 GeV in the hard process but got the rest from initial-state radiation would be missed. Not only therefore would cross sections come out wrong, but so might the typical event shapes. In the end, it is therefore necessary to find some reasonable compromise, by starting the generation at 30 GeV, say, if one knows that only rarely do events below this value fluctuate up to 50 GeV. Of course, most events will therefore not contain a $\mathrm{Z}$ above 50 GeV, and one will have to live with some inefficiency. It is not uncommon that only one event out of ten can be used, and occasionally it can be even worse. If it is difficult to set kinematics, it is often easier to set the flavour content of a process. In a Higgs study, one might wish, for example, to consider the decay $\H^0 \rightarrow \mathrm{Z}^0 \mathrm{Z}^0$, with each $\mathrm{Z}^0 \rightarrow \e^+\e^-$ or $\mu^+ \mu^-$. It is therefore necessary to inhibit all other $\H^0$ and $\mathrm{Z}^0$ decay channels, and also to adjust cross sections to take into account this change, all of which is fairly straightforward. However, if one wanted to consider instead the decay $\mathrm{Z}^0 \rightarrow \c \overline{\mathrm{c}}$, with a $\mathrm{D}$ meson producing a lepton, not only would there then be the problem of different leptonic branching ratios for different $\mathrm{D}$:s (which means that fragmentation and decay treatments would no longer decouple), but also that of additional $\c \overline{\mathrm{c}}$ pair production in parton-shower evolution, at a rate that is unknown beforehand. In practice, it is therefore impossible to force $\mathrm{D}$ decay modes in a consistent manner. \clearpage \section{Monte Carlo Techniques} Quantum mechanics introduces a concept of randomness in the behaviour of physical processes. The virtue of event generators is that this randomness can be simulated by the use of Monte Carlo techniques. In the process, the program authors have to use some ingenuity to find the most efficient way to simulate an assumed probability distribution. A detailed description of possible techniques would carry us too far, but in this section some of the most frequently used approaches are presented, since they will appear in discussions in subsequent sections. Further examples may be found e.g. in \cite{Jam80}. First of all one assumes the existence of a random number generator. This is a (Fortran) function which, each time it is called, returns a number $R$ in the range between 0 and 1, such that the inclusive distribution of numbers $R$ is flat in the range, and such that different numbers $R$ are uncorrelated. The random number generator that comes with {\tsc{Jetset}} is described at the end of this section, and we defer the discussion until then. \subsection{Selection From a Distribution} \label{ss:MCdistsel} The situation that is probably most common is that we know a function $f(x)$ which is non-negative in the allowed $x$ range $x_{\mathrm{min}} \leq x \leq x_{\mathrm{max}}$. We want to select an $x$ `at random' so that the probability for a given $x$ is proportional to $f(x)$. Here $f(x)$ might be a fragmentation function, a differential cross section, or any of a number of distributions. One does not have to assume that the integral of $f(x)$ is explicitly normalized to unity: by the Monte Carlo procedure of picking exactly one accepted $x$ value, normalization is implicit in the final result. Sometimes the integral of $f(x)$ does carry a physics content of its own, as part of an overall weight factor we want to keep track of. Consider, for instance, the case when $x$ represents one or several phase-space variables and $f(x)$ a differential cross section; here the integral has a meaning of total cross section for the process studied. The task of a Monte Carlo is then, on the one hand, to generate events one at a time, and, on the other hand, to estimate the total cross section. The discussion of this important example is deferred to section \ref{ss:PYTcrosscalc}. If it is possible to find a primitive function $F(x)$ which has a known inverse $F^{-1}(x)$, an $x$ can be found as follows (method 1): \begin{eqnarray} & \displaystyle{ \int_{x_{\mathrm{min}}}^{x} f(x) \, \d x = R \int_{x_{\mathrm{min}}}^{x_{\mathrm{max}}} f(x) \, \d x } & \nonumber \\[1mm] \Longrightarrow & x = F^{-1}(F(x_{\mathrm{min}}) + R(F(x_{\mathrm{max}}) - F(x_{\mathrm{min}}))) ~. & \end{eqnarray} The statement of the first line is that a fraction $R$ of the total area under $f(x)$ should be to the left of $x$. However, seldom are functions of interest so nice that the method above works. It is therefore necessary to use more complicated schemes. Special tricks can sometimes be found. Consider e.g. the generation of a Gaussian $f(x) = \exp(-x^2)$. This function is not integrable, but if we combine it with the same Gaussian distribution of a second variable $y$, it is possible to transform to polar coordinates \begin{equation} f(x) \, \d x \, f(y) \, \d y = \exp(-x^2-y^2) \, \d x \, \d y = r \exp(-r^2) \, \d r \, \d \varphi ~, \end{equation} and now the $r$ and $\varphi$ distributions may be easily generated and recombined to yield $x$. At the same time we get a second number $y$, which can also be used. For the generation of transverse momenta in fragmentation, this is very convenient, since in fact we want to assign two transverse degrees of freedom. If the maximum of $f(x)$ is known, $f(x) \leq f_{\mathrm{max}}$ in the $x$ range considered, a hit-or-miss method will always yield the correct answer (method 2): \begin{Enumerate} \item select an $x$ with even probability in the allowed range, i.e. $x = x_{\mathrm{min}} + R(x_{\mathrm{max}} - x_{\mathrm{min}})$; \item compare a (new) $R$ with the ratio $f(x)/f_{\mathrm{max}}$; if $f(x)/f_{\mathrm{max}} \le R$, then reject the $x$ value and return to point 1 for a new try; \item otherwise the most recent $x$ value is retained as final answer. \end{Enumerate} The probability that $f(x)/f_{\mathrm{max}} > R$ is proportional to $f(x)$; hence the correct distribution of retained $x$ values. The efficiency of this method, i.e. the average probability that an $x$ will be retained, is $(\int \, f(x) \, \d x) / (f_{\mathrm{max}}(x_{\mathrm{max}} - x_{\mathrm{min}}))$. The method is acceptable if this number is not too low, i.e. if $f(x)$ does not fluctuate too wildly. Very often $f(x)$ does have narrow spikes, and it may not even be possible to define an $f_{\mathrm{max}}$. An example of the former phenomenon is a function with a singularity just outside the allowed region, an example of the latter an integrable singularity just at the $x_{\mathrm{min}}$ and/or $x_{\mathrm{max}}$ borders. Variable transformations may then be used to make a function smoother. Thus a function $f(x)$ which blows up as $1/x$ for $x \rightarrow 0$, with an $x_{\mathrm{min}}$ close to 0, would instead be roughly constant if transformed to the variable $y = \ln x$. The variable transformation strategy may be seen as a combination of methods 1 and 2, as follows. Assume the existence of a function $g(x)$, with $f(x) \leq g(x)$ over the $x$ range of interest. Here $g(x)$ is picked to be a `simple' function, such that the primitive function $G(x)$ and its inverse $G^{-1}(x)$ are known. Then (method 3): \begin{Enumerate} \item select an $x$ according to the distribution $g(x)$, using method 1; \item compare a (new) $R$ with the ratio $f(x)/g(x)$; if $f(x)/g(x) \le R$, then reject the $x$ value and return to point 1 for a new try; \item otherwise the most recent $x$ value is retained as final answer. \end{Enumerate} This works, since the first step will select $x$ with a probability $g(x) \, \d x = \d G(x)$ and the second retain this choice with probability $f(x)/g(x)$. The total probability to pick a value $x$ is then just the product of the two, i.e. $f(x) \, \d x$. If $f(x)$ has several spikes, method 3 may work for each spike separately, but it may not be possible to find a $g(x)$ that covers all of them at the same time, and which still has an invertible primitive function. However, assume that we can find a function $g(x) = \sum_i g_i(x)$, such that $f(x) \leq g(x)$ over the $x$ range considered, and such that the functions $g_i(x)$ each are non-negative and simple, in the sense that we can find primitive functions and their inverses. In that case (method 4): \begin{Enumerate} \item select an $i$ at random, with relative probability given by the integrals \begin{equation} \int_{x_{\mathrm{min}}}^{x_{\mathrm{max}}} g_i(x) \, \d x = G_i(x_{\mathrm{max}}) - G_i(x_{\mathrm{min}}) ~; \nonumber \end{equation} \item for the $i$ selected, use method 1 to find an $x$, i.e. \begin{equation} x = G_i^{-1}(G_i(x_{\mathrm{min}}) + R(G_i(x_{\mathrm{max}})-G_i(x_{\mathrm{min}}))) ~; \nonumber \end{equation} \item compare a (new) $R$ with the ratio $f(x)/g(x)$; if $f(x)/g(x) \le R$, then reject the $x$ value and return to point 1 for a new try; \item otherwise the most recent $x$ value is retained as final answer. \end{Enumerate} This is just a trivial extension of method 3, where steps 1 and 2 ensure that, on the average, each $x$ value picked there is distributed according to $g(x)$: the first step picks $i$ with relative probability $\int g_i(x) \, \d x$, the second $x$ with absolute probability $g_i(x) / \int g_i(x) \, \d x$ (this is one place where one must remember to do normalization correctly); the product of the two is therefore $g_i(x)$ and the sum over all $i$ gives back $g(x)$. We have now arrived at an approach that is sufficiently powerful for a large selection of problems. In general, for a function $f(x)$ which is known to have sharp peaks in a few different places, the generic behaviour at each peak separately may be covered by one or a few simple functions $g_i(x)$, to which one adds a few more $g_i(x)$ to cover the basic behaviour away from the peaks. By a suitable selection of the relative strengths of the different $g_i$'s, it is possible to find a function $g(x)$ that matches well the general behaviour of $f(x)$, and thus achieve a reasonable Monte Carlo efficiency. The major additional complication is when $x$ is a multidimensional variable. Usually the problem is not so much $f(x)$ itself, but rather that the phase-space boundaries may be very complicated. If the boundaries factorize it is possible to pick phase-space points restricted to the desired region. Otherwise the region may have to be inscribed in a hyper-rectangle, with points picked within the whole hyper-rectangle but only retained if they are inside the allowed region. This may lead to a significant loss in efficiency. Variable transformations may often make the allowed region easier to handle. There are two main methods to handle several dimensions, each with its set of variations. The first method is based on a factorized ansatz, i.e. one attempts to find a function $g(\mbf{x})$ which is everywhere larger than $f(\mbf{x})$, and which can be factorized into $g(\mbf{x}) = g^{(1)}(x_1) \, g^{(2)}(x_2) \cdots g^{(n)}(x_n)$, where $\mbf{x} = (x_1,x_2,\ldots,x_n)$. Here each $g^{(j)}(x_j)$ may in its turn be a sum of functions $g^{(j)}_i$, as in method 4 above. First, each $x_j$ is selected independently, and afterwards the ratio $f(\mbf{x})/g(\mbf{x})$ is used to determine whether to retain the point. The second method is useful if the boundaries of the allowed region can be written in a form where the maximum range of $x_1$ is known, the allowed range of $x_2$ only depends on $x_1$, that of $x_3$ only on $x_1$ and $x_2$, and so on until $x_n$, whose range may depend on all the preceding variables. In that case it may be possible to find a function $g(\mbf{x})$ that can be integrated over $x_2$ through $x_n$ to yield a simple function of $x_1$, according to which $x_1$ is selected. Having done that, $x_2$ is selected according to a distribution which now depends on $x_1$, but with $x_3$ through $x_n$ integrated over. In particular, the allowed range for $x_2$ is known. The procedure is continued until $x_n$ is reached, where now the function depends on all the preceding $x_j$ values. In the end, the ratio $f(\mbf{x})/g(\mbf{x})$ is again used to determine whether to retain the point. \subsection{The Veto Algorithm} \label{ss:vetoalg} The `radioactive decay' type of problems is very common, in particular in parton showers, but it is also used, e.g. in the multiple interactions description in {\tsc{Pythia}}. In this kind of problems there is one variable $t$, which may be thought of as giving a kind of time axis along which different events are ordered. The probability that `something will happen' (a nucleus decay, a parton branch) at time $t$ is described by a function $f(t)$, which is non-negative in the range of $t$ values to be studied. However, this na\"{\i}ve probability is modified by the additional requirement that something can only happen at time $t$ if it did not happen at earlier times $t' < t$. (The original nucleus cannot decay once again if it already did decay; possibly the decay products may decay in their turn, but that is another question.) The probability that nothing has happened by time $t$ is expressed by the function ${\cal N}(t)$ and the differential probability that something happens at time $t$ by ${\cal P}(t)$. The basic equation then is \begin{equation} {\cal P}(t) = - \frac{\d {\cal N}}{\d t} = f(t) \, {\cal N}(t) ~. \end{equation} For simplicity, we shall assume that the process starts at time $t = 0$, with ${\cal N}(0) = 1$. The above equation can be solved easily if one notes that $\d {\cal N} / {\cal N} = \d \ln {\cal N}$: \begin{equation} {\cal N}(t) = {\cal N}(0) \exp \left\{ - \int_0^t f(t') \, \d t' \right\} = \exp \left\{ - \int_0^t f(t') \, \d t' \right\} ~, \end{equation} and thus \begin{equation} {\cal P}(t) = f(t) \exp \left\{ - \int_0^t f(t') \, \d t' \right\} ~. \label{mc:Pveto} \end{equation} With $f(t) = c$ this is nothing but the textbook formulae for radioactive decay. In particular, at small times the correct decay probability, ${\cal P}(t)$, agrees well with the input one, $f(t)$, since the exponential factor is close to unity there. At larger $t$, the exponential gives a dampening which ensures that the integral of ${\cal P}(t)$ never can exceed unity, even if the integral of $f(t)$ does. The exponential can be seen as the probability that nothing happens between the original time 0 and the final time $t$. In the parton-shower language, this is (almost) the so-called Sudakov form factor. If $f(t)$ has a primitive function with a known inverse, it is easy to select $t$ values correctly: \begin{equation} \int_0^t {\cal P}(t') \, \d t' = {\cal N}(0) - {\cal N}(t) = 1 - \exp \left\{ - \int_0^t f(t') \, \d t' \right\} = 1 - R ~, \end{equation} which has the solution \begin{equation} F(0) - F(t) = \ln R ~~~ \Longrightarrow ~~~ t = F^{-1}(F(0) - \ln R) ~. \end{equation} If $f(t)$ is not sufficiently nice, one may again try to find a better function $g(t)$, with $f(t) \leq g(t)$ for all $t \geq 0$. However to use method 3 with this $g(t)$ would not work, since the method would not correctly take into account the effects of the exponential term in ${\cal P}(t)$. Instead one may use the so-called veto algorithm: \begin{Enumerate} \item start with $i = 0$ and $t_0 = 0$; \item add 1 to $i$ and select $t_i = G^{-1}(G(t_{i-1}) - \ln R)$, i.e. according to $g(t)$, but with the constraint that $t_i > t_{i-1}$, \item compare a (new) $R$ with the ratio $f(t_i)/g(t_i)$; if $f(t_i)/g(t_i) \le R$, then return to point 2 for a new try; \item otherwise $t_{i}$ is retained as final answer. \end{Enumerate} It may not be apparent why this works. Consider, however, the various ways in which one can select a specific time $t$. The probability that the first try works, $t = t_1$, i.e. that no intermediate $t$ values need be rejected, is given by \begin{equation} {\cal P}_0 (t) = \exp \left\{ - \int_0^t g(t') \, \d t' \right\} \, g(t) \, \frac{f(t)}{g(t)} = f(t) \exp \left\{ - \int_0^t g(t') \, \d t' \right\} ~, \end{equation} where the exponential times $g(t)$ comes from eq.~(\ref{mc:Pveto}) applied to $g$, and the ratio $f(t)/g(t)$ is the probability that $t$ is accepted. Now consider the case where one intermediate time $t_1$ is rejected and $t = t_2$ is only accepted in the second step. This gives \begin{equation} {\cal P}_1 (t) = \int_0^t \d t_1 \exp \left\{ - \int_0^{t_1} g(t') \, \d t' \right\} g(t_1) \left[ 1 - \frac{f(t_1)}{g(t_1)} \right] \exp \left\{ - \int_{t_1}^t g(t') \, \d t' \right\} g(t) \, \frac{f(t)}{g(t)} ~, \end{equation} where the first exponential times $g(t_1)$ gives the probability that $t_1$ is first selected, the square brackets the probability that $t_1$ is subsequently rejected, the following piece the probability that $t = t_2$ is selected when starting from $t_1$, and the final factor that $t$ is retained. The whole is to be integrated over all possible intermediate times $t_1$. The exponentials together give an integral over the range from 0 to $t$, just as in ${\cal P}_0$, and the factor for the final step being accepted is also the same, so therefore one finds that \begin{equation} {\cal P}_1 (t) = {\cal P}_0 (t) \int_0^t \d t_1 \left[ g(t_1) - f(t_1) \right] ~. \end{equation} This generalizes. In ${\cal P}_2$ one has to consider two intermediate times, $0 \leq t_1 \leq t_2 \leq t_3 = t$, and so \begin{eqnarray} {\cal P}_2 (t) & = & {\cal P}_0 (t) \int_0^t \d t_1 \left[ g(t_1) - f(t_1) \right] \int_{t_1}^t \d t_2 \left[ g(t_2) - f(t_2) \right] \nonumber \\ & = & {\cal P}_0 (t) \frac{1}{2} \left( \int_0^t \left[ g(t') - f(t') \right] \d t' \right)^2 ~. \end{eqnarray} The last equality is most easily seen if one also considers the alternative region $0 \leq t_2 \leq t_1 \leq t$, where the r\^oles of $t_1$ and $t_2$ have just been interchanged, and the integral therefore has the same value as in the region considered. Adding the two regions, however, the integrals over $t_1$ and $t_2$ decouple, and become equal. In general, for ${\cal P}_i$, the $i$ intermediate times can be ordered in $i!$ different ways. Therefore the total probability to accept $t$, in any step, is \begin{eqnarray} {\cal P} (t) & = & \displaystyle{ \sum_{i=0}^{\infty} {\cal P}_i (t) = {\cal P}_0 (t) \sum_{i=0}^{\infty} \frac{1}{i!} \left( \int_0^t \left[ g(t') - f(t') \right] \d t' \right)^i } \nonumber \\ & = & \displaystyle{ f(t) \exp \left\{ - \int_0^t g(t') \, \d t' \right\} \exp \left\{ \int_0^t \left[ g(t') - f(t') \right] \d t' \right\} } \nonumber \\ & = & \displaystyle{ f(t) \exp \left\{ - \int_0^t f(t') \, \d t' \right\} ~, } \end{eqnarray} which is the desired answer. If the process is to be stopped at some scale $t_{\mathrm{max}}$, i.e. if one would like to remain with a fraction ${\cal N}(t_{\mathrm{max}})$ of events where nothing happens at all, this is easy to include in the veto algorithm: just iterate upwards in $t$ at usual, but stop the process if no allowed branching is found before $t_{\mathrm{max}}$. Usually $f(t)$ is a function also of additional variables $x$. The methods of the preceding subsection are easy to generalize if one can find a suitable function $g(t,x)$ with $f(t,x) \leq g(t,x)$. The $g(t)$ used in the veto algorithm is the integral of $g(t,x)$ over $x$. Each time a $t_i$ has been selected also an $x_i$ is picked, according to $g(t_i,x) \, dx$, and the $(t,x)$ point is accepted with probability $f(t_i,x_i)/g(t_i,x_i)$. \subsection{The Random Number Generator} The construction of a good, portable (pseudo)random generator is not a trivial task. Therefore {\tsc{Jetset}} has traditionally stayed away from that area, and just provided the routine \ttt{RLU} as an interface, which the user could modify to call on an existing routine, implemented on the actual machine being used. In recent years, progress has been made in constructing portable generators with large periods and other good properties; see the review \cite{Jam90}. Therefore the current version contains a random number generator based on the algorithm proposed by Marsaglia, Zaman and Tsang \cite{Mar90}. This routine should work on any machine with a mantissa of at least 24 digits, i.e. all common 32-bit (or more) computers. Given the same initial state, the sequence will also be identical on different machines. This need not mean that the same sequence of events will be generated on an IBM and a VAX, say, since the different treatments of roundoff errors in numerical operations will lead to slightly different real numbers being tested against these random numbers in IF statements. Also code optimization may lead to a divergence. Apart from nomenclature issues, and the coding of \ttt{RLU} as a function rather than a subroutine, the only difference between the {\tsc{Jetset}} code and the code given in \cite{Jam90} is that slightly different algorithms are used to ensure that the random number is not equal to 0 or 1 within the machine precision. The generator has a period of over $10^{43}$, and the possibility to obtain almost $10^9$ different and disjoint subsequences, selected by giving an initial integer number. The price to be paid for the long period is that the state of the generator at a given moment cannot be described by a single integer, but requires about 100 words. Some of these are real numbers, and are thus not correctly represented in decimal form. The normal procedure, which makes it possible to restart the generation from a seed value written to the run output, is therefore not convenient. The CERN library implementation keeps track of the number of random numbers generated since the start. With this value saved, in a subsequent run the random generator can be asked to skip ahead the corresponding number of random numbers. {\tsc{Jetset}} is a heavy user of random numbers, however: typically 30\% of the full run time is spent on random number generation. Of this, half is overhead coming from the function call administration, but the other half is truly related to the speed of the algorithm. Therefore a skipping ahead would take place with 15\% of the time cost of the original run, i.e. an uncomfortably high figure. Instead a different solution is chosen here. Two special routines are provided for writing and reading the state of the random number generator (plus some initialization information) on a sequential file, in a machine-dependent internal representation. The file used for this purpose has to be specified by you, and opened for read and write. A state is written as a single record, in free format. It is possible to write an arbitrary number of states on a file, and a record can be overwritten, if so desired. The event generation loop might then look something like: \begin{Enumerate} \item save the state of the generator on file (using flag set in point 3 below), \item generate an event, \item study the event for errors or other reasons why to regenerate it later; set flag to overwrite previous generator state if no errors, otherwise set flag to create new record; \item loop back to point 1. \end{Enumerate} With this procedure, the file will contain the state before each of the problematical events. An alternative approach might be to save the state every 100 events or so. If the events are subsequently processed through a detector simulation, you may have to save also other sets of seeds, naturally. In addition to the service routines, the common block which contains the state of the generator is available for manipulation, if you so desire. In particular, the initial seed value is by default 19780503, i.e. different from the Marsaglia/CERN default 54217137. It is possible to change this value before any random numbers have been generated, or to force reinitialization in mid-run with any desired new seed. Inside {\tsc{Jetset/Pythia}}, some initialization may take place in connection with the very first event generated in a run, so sometimes it may be necessary to generate one ordinary event before reading in a saved state to generate an interesting event. In the current {\tsc{Pythia}} version, some of the multiple interaction machinery options contain an element of learning, which means that the event sequence may be broken. It should be noted that, of course, the appearance of a random number generator package inside {\tsc{Jetset}} does in no way preclude the use of other routines. You can easily revert to the old approach, where \ttt{RLU} is nothing but an interface to an arbitrary external random number generator; e.g. to call a routine \ttt{RNDM} all you need to have is \begin{verbatim} FUNCTION RLU(IDUMMY) 100 RLU=RNDM(IDUMMY) IF(RLU.LE.0..OR.RLU.GE.1.) GOTO 100 RETURN END \end{verbatim} The random generator subpackage consists of the following components. \drawbox{R = RLU(IDUMMY)}\label{p:RLU} \begin{entry} \itemc{Purpose:} to generate a (pseudo)random number \ttt{R} uniformly in the range 0$<$\ttt{R}$<$1, i.e. excluding the endpoints. \iteme{IDUMMY :} dummy input argument; normally 0. \end{entry} \drawbox{CALL RLUGET(LFN,MOVE)}\label{p:RLUGET} \begin{entry} \itemc{Purpose:} to dump the current state of the random number generator on a separate file, using internal representation for real and integer numbers. To be precise, the full contents of the \ttt{LUDATR} common block are written on the file, with the exception of \ttt{MRLU(6)}. \iteme{LFN :} (logical file number) the file number to which the state is dumped. You must associate this number with a true file (with a machine-dependent name), and see to it that this file is open for write. \iteme{MOVE :} choice of adding a new record to the file or overwriting old record(s). Normally only options 0 or $-$1 should be used. \begin{subentry} \iteme{= 0 (or > 0) :} add a new record to the end of the file. \iteme{= -1 :} overwrite the last record with a new one (i.e. do one \ttt{BACKSPACE} before the new write). \iteme{= $-n$ :} back up $n$ records before writing the new record. The records following after the new one are lost, i.e. the last $n$ old records are lost and one new added. \end{subentry} \end{entry} \drawbox{CALL RLUSET(LFN,MOVE)}\label{p:RLUSET} \begin{entry} \itemc{Purpose:} to read in a state for the random number generator, from which the subsequent generation can proceed. The state must previously have been saved by a \ttt{RLUGET} call. Again the full contents of the \ttt{LUDATR} common block are read, with the exception of \ttt{MRLU(6)}. \iteme{LFN :} (logical file number) the file number from which the state is read. You must associate this number with a true file previously written with a \ttt{RLUGET} call, and see to it that this file is open for read. \iteme{MOVE :} positioning in file before a record is read. With zero value, records are read one after the other for each new call, while non-zero values may be used to navigate back and forth, and e.g. return to the same initial state several times. \begin{subentry} \iteme{= 0 :} read the next record. \iteme{= $+n$ :} skip ahead $n$ records before reading the record that sets the state of the random number generator. \iteme{= $-n$ :} back up $n$ records before reading the record that sets the state of the random number generator. \end{subentry} \end{entry} \drawbox{COMMON/LUDATR/MRLU(6),RRLU(100)}\label{p:LUDATR} \begin{entry} \itemc{Purpose:} to contain the state of the random number generator at any moment (for communication between \ttt{RLU}, \ttt{RLUGET} and \ttt{RLUSET}), and also to provide the user with the possibility to initialize different random number sequences, and to know how many numbers have been generated. \iteme{MRLU(1) :}\label{p:MRLU} (D=19780503) the integer number that specifies which of the possible subsequences will be initialized in the next \ttt{RLU} call for which \ttt{MRLU(2)=0}. Allowed values are 0$\leq$\ttt{MRLU(1)}$\leq$900\,000\,000, the original Marsaglia (and CERN library) seed is 54217137. The \ttt{MRLU(1)} value is not changed by any of the {\tsc{Jetset}} routines. \iteme{MRLU(2) :} (D=0) initialization flag, put to 1 in the first \ttt{RLU} call of run. A reinitialization of the random number generator can be made in mid-run by resetting \ttt{MRLU(2)} to 0 by hand. In addition, any time the counter \ttt{MRLU(3)} reaches 1000000000, it is reset to 0 and \ttt{MRLU(2)} is increased by 1. \iteme{MRLU(3) :} (D=0) counter for the number of random numbers generated from the beginning of the run. To avoid overflow when very many numbers are generated, \ttt{MRLU(2)} is used as described above. \iteme{MRLU(4), MRLU(5) :} \ttt{I97} and \ttt{J97} of the CERN library implementation; part of the state of the generator. \iteme{MRLU(6) :} (D=0) current position, i.e. how many records after beginning, in the file; used by \ttt{RLUGET} and \ttt{RLUSET}. \iteme{RRLU(1) - RRLU(97) :}\label{p:RRLU} the \ttt{U} array of the CERN library implementation; part of the state of the generator. \iteme{RRLU(98) - RRLU(100) :} \ttt{C}, \ttt{CD} and \ttt{CM} of the CERN library implementation; the first part of the state of the generator, the latter two constants calculated at initialization. \end{entry} \clearpage \section{The Event Record} The event record is the central repository for information about the particles produced in the current event: flavours, momenta, event history, and production vertices. It plays a very central r\^ole: without a proper understanding of what the record is and how information is stored, it is meaningless to try to use either {\tsc{Jetset}} or {\tsc{Pythia}}. The record is stored in the common block \ttt{LUJETS}. Almost all the routines thatthe user calls can be viewed as performing some action on the record: fill a new event, let partons fragment or particles decay, boost it, list it, find clusters, etc. In this section we will first describe the KF flavour code, subsequently the \ttt{LUJETS} common block, and then give a few comments about the r\^ole of the event record in the programs. To ease the interfacing of different event generators, a \ttt{HEPEVT} standard common block structure for the event record has been agreed on. For historical reasons the standard common blocks are not directly used in {\tsc{Jetset}}, but a conversion routine comes with the program, and is described at the end of this section. \subsection{Particle Codes} \label{ss:codes} The new particle code now adopted by the Particle Data Group \cite{PDG88,PDG92} is used consistently throughout the program, and is referred to as the KF particle code. This code you have to be thoroughly familiar with. It is described below. Note that a few inconsistencies between the KF and the PDG codes are known, which stem from differences of interpretation of the rules agreed on when developing the standard. These rules form the basis of the PDG tables and (independently) of the {\tsc{Jetset}} tables. (Of course, my private opinion is that I follow the original agreement, and the PDG deviate from it.) Hopefully, this should have few practical consequences, since only rarely-produced particles are affected. Anyway, here is a list of the known discrepancies: \begin{Enumerate} \item The PDG has not allowed for the existence of an $\eta_{\b}$, which in {\tsc{Jetset}} is included with code 551. This code is reserved for $\chi_{0 \b}$ by the PDG, a particle which appears as 10551 in {\tsc{Jetset}}. (We agree to have $\eta_{\c}$ as 441, which illustrates the basic difference: I use the additional recurrence figure to refer to a whole multiplet, whether all particles of that multiplet have been found or not; the PDG, on the other hand, does not reserve space for particles which we know should be there but have not yet been discovered, which means that members of a multiplet need not go together.) \item The PDG has not allowed for the existence of an $\mathrm{h}_{1 \c}$, which in {\tsc{Jetset}} is represented by 10443. Therefore $\chi_{1 \c}$ is the PDG code 10443 but {\tsc{Jetset}} code 20443. Further $\psi'$ is either 20443 or 30443, and $\Upsilon' = \Upsilon(2S)$ either 20553 or 30553. (Comment as for point 1.) \item Different conventions for spin $1/2$ baryons with one heavy flavour (charm, bottom, top), one strange flavour, and one light ($\u$ or $\d$). Here two states exist, e.g. $\Xi_{\c}^+$ and $\Xi'^+_{\c}$, both with flavour content $\c\mathrm{s}\u$. By analogy with the $\Lambda^0$--$\Sigma$ pair, {\tsc{Jetset}} uses the decreasing order of flavour content for the heavier state and inversed order of the two lighter flavours for the lighter state, while the PDG tables use the opposite convention. Thus in {\tsc{Jetset}} $\Xi_{\c}^+$ is 4232 and $\Xi'^+_{\c}$ 4322, while in PDG it is the other way around. \end{Enumerate} There are no plans to change the {\tsc{Jetset}} rules to agree with the PDG ones in either of the cases above. The KF code is not convenient for a direct storing of masses, decay data, or other particle properties, since the KF codes are so spread out. Instead a compressed code KC between 1 and 500 is used here, where the most frequently used particles have a separate code, but many heavy-flavour hadrons are lumped together in groups. Normally this code is only used at very specific places in the program, not visible to the user. If need be, the correspondence can always be obtained by using the function \ttt{LUCOMP}, \mbox{\ttt{KC = LUCOMP(KF)}}. It is therefore not intended that you should ever need to know any KC codes at all. It may be useful to know, however, that for codes smaller than 80, KF and KC agree. The particle names printed in the tables in this section correspond to the ones obtained with the routine \ttt{LUNAME}, which is used extensively, e.g. in \ttt{LULIST}. Greek characters are spelt out in full, with a capital first letter to correspond to a capital Greek letter. Generically the name of a particle is made up of the following pieces: \begin{Enumerate} \item The basic root name. This includes a * for most spin 1 ($L = 0$) mesons and spin $3/2$ baryons, and a $'$ for some spin $1/2$ baryons (where there are two states to be distinguished, cf. $\Lambda$--$\Sigma^0$). The rules for heavy baryon naming are in accordance with the 1986 Particle Data Group conventions \cite{PDG86}. For mesons with one unit of orbital angular momentum, K (D, B, \ldots) is used for quark-spin 0 and K* (D*, B*, \ldots) for quark-spin 1 mesons; the convention for `*' may here deviate slightly from the one used by the PDG. \item Any lower indices, separated from the root by a \_. For heavy hadrons, this is the additional heavy-flavour content not inherent in the root itself. For a diquark, it is the spin. \item The character $\sim$ (alternatively bar, see \ttt{MSTU(15)}) for an antiparticle, wherever the distinction between particle and antiparticle is not inherent in the charge information. \item Charge information: $++$, $+$, $0$, $-$, or $--$. Charge is not given for quarks or diquarks. Some neutral particles which are customarily given without a 0 also here lack it, such as neutrinos, $\mathrm{g}$, $\gamma$, and flavour-diagonal mesons other than $\pi^0$ and $\rho^0$. Note that charge is included both for the proton and the neutron. While non-standard, it is helpful in avoiding misunderstandings when looking at an event listing. \end{Enumerate} Below follows a list of KF particle codes. The list is not complete; a more extensive one may be obtained with \ttt{CALL LULIST(11)}. Particles are grouped together, and the basic rules are described for each group. Whenever a distinct antiparticle exists, it is given the same KF code with a minus sign (whereas KC codes are always positive). \begin{Enumerate} \item Quarks and leptons, Table \ref{t:codeone}. \\ This group contains the basic building blocks of matter, arranged according to family, with the lower member of weak isodoublets also having the smaller code (thus $\d$ precedes $\u$, contrary to the ordering in previous {\tsc{Jetset}} versions). A fourth generation is included for future reference. The quark codes are used as building blocks for the diquark, meson and baryon codes below. \begin{table}[ptb] \captive{Quark and lepton codes. \protect\label{t:codeone} } \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c||c|c|c|@{\protect\rule{0mm}{\tablinsep}}} \hline KF & Name & Printed & KF & Name & Printed \\ \hline 1 & $\d$ & \ttt{d} & 11 & $\mathrm{e}^-$ & \ttt{e-} \\ 2 & $\u$ & \ttt{u} & 12 & $\nu_{\mathrm{e}}$ & \ttt{nu\_e} \\ 3 & $\mathrm{s}$ & \ttt{s} & 13 & $\mu^-$ & \ttt{mu-} \\ 4 & $\c$ & \ttt{c} & 14 & $\nu_{\mu}$ & \ttt{nu\_mu} \\ 5 & $\b$ & \ttt{b} & 15 & $\tau^-$ & \ttt{tau-} \\ 6 & $\t$ & \ttt{t} & 16 & $\nu_{\tau}$ & \ttt{nu\_tau} \\ 7 & $\mathrm{l}$ & \ttt{l} & 17 & $\chi^-$ & \ttt{chi-} \\ 8 & $\mathrm{h}$ & \ttt{h} & 18 & $\nu_{\chi}$ & \ttt{nu\_chi} \\ 9 & & & 19 & & \\ 10 & & & 20 & & \\ \hline \end{tabular} \end{center} \end{table} \item Gauge bosons and other fundamental bosons, Table \ref{t:codetwo}. \\ This group includes all the gauge and Higgs bosons of the standard model, as well as some of the bosons appearing in various extensions of it. The latter are not covered by the standard PDG codes. They correspond to one extra {\bf U(1)} group and one extra {\bf SU(2)} one, a further Higgs doublet, a (scalar, colour octet) techni-$\eta$, a (scalar) leptoquark $\L_{\mathrm{Q}}$, and a horizontal gauge boson $\mathrm{R}$ (coupling between families). Additionally, we here include the pomeron $\mathrm{I}\!\mathrm{P}$ and reggeon $\mathrm{I}\!\mathrm{R}$ `particles', which are important e.g. in the description of diffractive scattering, but have no obvious position anywhere in the classification scheme. \begin{table}[ptb] \captive{Gauge boson and other fundamental boson codes. \protect\label{t:codetwo} } \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c||c|c|c|@{\protect\rule{0mm}{\tablinsep}}} \hline KF & Name & Printed & KF & Name & Printed \\ \hline 21 & $\mathrm{g}$ & \ttt{g} & 31 & & \\ 22 & $\gamma$ & \ttt{gamma} & 32 & $\mathrm{Z}'^0$ & \ttt{Z'0} \\ 23 & $\mathrm{Z}^0$ & \ttt{Z0} & 33 & $\mathrm{Z}''^0$ & \ttt{Z"0} \\ 24 & $\mathrm{W}^+$ & \ttt{W+} & 34 & $\mathrm{W}'^+$ & \ttt{W'+} \\ 25 & $\H^0$ & \ttt{H0} & 35 & $\H'^0$ & \ttt{H'0} \\ 26 & & & 36 & $\mathrm{A}^0$ & \ttt{A0} \\ 27 & & & 37 & $\H^+$ & \ttt{H+} \\ 28 & $\mathrm{I}\!\mathrm{R}$ & \ttt{reggeon} & 38 & $\eta_{\mrm{techni}}$ & \ttt{eta\_tech0} \\ 29 & $\mathrm{I}\!\mathrm{P}$ & \ttt{pomeron} & 39 & $\L_{\mathrm{Q}}$ & \ttt{LQ} \\ 30 & & & 40 & $\mathrm{R}^0$ & \ttt{R0} \\ \hline \end{tabular} \end{center} \end{table} \item Free space. \\ The positions 41--80 are currently unused. In the future, they might come to be used, e.g. for supersymmetric partners of the particles above, or for some other kind of new physics. At the moment, they are at your disposal. \item Various special codes, Table \ref{t:codefour}. \\ In a Monte Carlo, it is always necessary to have codes that do not correspond to any specific particle, but are used to lump together groups of similar particles for decay treatment, or to specify generic decay products. These codes, which again are non-standard, are found between numbers 81 and 100. Several are not found in the event record, and therefore properly belong only to the KC group of codes. \begin{table}[ptb] \captive{Various special codes. \protect\label{t:codefour} } \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c|@{\protect\rule{0mm}{\tablinsep}}} \hline KF & Printed & Meaning \\ \hline 81 & \ttt{specflav} & Spectator flavour; used in decay-product listings \\ 82 & \ttt{rndmflav} & A random $\u$, $\d$, or $\mathrm{s}$ flavour; possible decay product \\ 83 & \ttt{phasespa} & Simple isotropic phase-space decay \\ 84 & \ttt{c-hadron} & Information on decay of generic charm hadron \\ 85 & \ttt{b-hadron} & Information on decay of generic bottom hadron \\ 86 & \ttt{t-hadron} & Information on decay of generic top hadron \\ 87 & \ttt{l-hadron} & Information on decay of generic low hadron \\ 88 & \ttt{h-hadron} & Information on decay of generic high hadron \\ 89 & \ttt{Wvirt} & Off-mass-shell $\mathrm{W}$ in weak decays of $\t$, $\mathrm{l}$, $\mathrm{h}$ or $\chi$ \\ 90 & \ttt{diquark} & Generic code for diquark colour information \\ 91 & \ttt{cluster} & Parton system in cluster fragmentation \\ 92 & \ttt{string} & Parton system in string fragmentation \\ 93 & \ttt{indep.} & Parton system in independent fragmentation \\ 94 & \ttt{CMshower} & Four-momentum of time-like showering system \\ 95 & \ttt{SPHEaxis} & Event axis found with \ttt{LUSPHE} \\ 96 & \ttt{THRUaxis} & Event axis found with \ttt{LUTHRU} \\ 97 & \ttt{CLUSjet} & Jet (cluster) found with \ttt{LUCLUS} \\ 98 & \ttt{CELLjet} & Jet (cluster) found with \ttt{LUCELL} \\ 99 & \ttt{table} & Tabular output from \ttt{LUTABU} \\ 100 & & \\ \hline \end{tabular} \end{center} \end{table} \item Diquark codes, Table \ref{t:codefive}. \\ A diquark made up of a quark with code $i$ and another with code $j$, where $i \geq j$, and with total spin $s$, is given the code \begin{equation} \mrm{KF} = 1000 i + 100 j + 2s + 1 ~, \end{equation} i.e. the tens position is left empty (cf. the baryon code below). Some of the most frequently used codes are listed in the table. All the lowest-lying spin 0 and 1 diquarks are included in the program. The corresponding KC code is 90, and it is mainly used to store colour charge. \begin{table}[ptb] \captive{Diquark codes. \protect\label{t:codefive} } \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c||c|c|c|@{\protect\rule{0mm}{\tablinsep}}} \hline KF & Name & Printed & KF & Name & Printed \\ \hline & & & 1103 & $\d\d_1$ & \ttt{dd\_1} \\ 2101 & $\u\d_0$ & \ttt{ud\_0} & 2103 & $\u\d_1$ & \ttt{ud\_1} \\ & & & 2203 & $\u\u_1$ & \ttt{uu\_1} \\ 3101 & $\mathrm{s}\d_0$ & \ttt{sd\_0} & 3103 & $\mathrm{s}\d_1$ & \ttt{sd\_1} \\ 3201 & $\mathrm{s}\u_0$ & \ttt{su\_0} & 3203 & $\mathrm{s}\u_1$ & \ttt{su\_1} \\ & & & 3303 & $\mathrm{s}\s_1$ & \ttt{ss\_1} \\ \hline \end{tabular} \end{center} \end{table} \item Meson codes, Tables \ref{t:codesixa} and \ref{t:codesixb}. \\ A meson made up of a quark with code $i$ and an antiquark with code $-j$, $j \neq i$, and with total spin $s$, is given the code \begin{equation} \mrm{KF} = \left\{ 100 \max(i,j) + 10 \min(i,j) + 2s + 1 \right\} \, \mrm{sign}(i-j) \, (-1)^{\max(i,j)} ~. \end{equation} Note the presence of an extra $-$ sign if the heaviest quark is a down-type one. This is in accordance with the particle--antiparticle distinction adopted in the 1986 Review of Particle Properties \cite{PDG86}. It means for example that a $\mathrm{B}$ meson contains a $\overline{\mathrm{b}}$ antiquark rather than a $\b$ quark. The flavour-diagonal states are arranged in order of ascending mass. The standard rule of having the last digit of the form $2s+1$ is broken for the $\mathrm{K}_{\mrm{S}}^0$--$\mathrm{K}_{\mrm{L}}^0$ system, where it is 0, and this convention should carry over to mixed states in the $\mathrm{B}$ meson system. For higher multiplets with the same spin, $\pm$10000, $\pm$20000, etc., are added to provide the extra distinction needed. Some of the most frequently used codes are given below. The full lowest-lying pseudoscalar and vector multiplets are included in the program, Table \ref{t:codesixa}. \begin{table}[ptb] \captive{Meson codes, part 1. \protect\label{t:codesixa} } \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c||c|c|c|@{\protect\rule{0mm}{\tablinsep}}} \hline KF & Name & Printed & KF & Name & Printed \\ \hline 211 & $\pi^+$ & \ttt{pi+} & 213 & $\rho^+$ & \ttt{rho+} \\ 311 & $\mathrm{K}^0$ & \ttt{K0} & 313 & $\mathrm{K}^{*0}$ & \ttt{K*0} \\ 321 & $\mathrm{K}^+$ & \ttt{K+} & 323 & $\mathrm{K}^{*+}$ & \ttt{K*+} \\ 411 & $\mathrm{D}^+$ & \ttt{D+} & 413 & $\mathrm{D}^{*+}$ & \ttt{D*+} \\ 421 & $\mathrm{D}^0$ & \ttt{D0} & 423 & $\mathrm{D}^{*0}$ & \ttt{D*0} \\ 431 & $\mathrm{D}_s^+$ & \ttt{D\_s+} & 433 & $\mathrm{D}_{\mathrm{s}}^{*+}$ & \ttt{D*\_s+} \\ 511 & $\mathrm{B}^0$ & \ttt{B0} & 513 & $\mathrm{B}^{*0}$ & \ttt{B*0} \\ 521 & $\mathrm{B}^+$ & \ttt{B+} & 523 & $\mathrm{B}^{*+}$ & \ttt{B*+} \\ 531 & $\mathrm{B}_s^0$ & \ttt{B\_s0} & 533 & $\mathrm{B}_{\mathrm{s}}^{*0}$ & \ttt{B*\_s0} \\ 541 & $\mathrm{B}_c^+$ & \ttt{B\_c+} & 543 & $\mathrm{B}_{\c}^{*+}$ & \ttt{B*\_c+} \\ 111 & $\pi^0$ & \ttt{pi0} & 113 & $\rho^0$ & \ttt{rho0} \\ 221 & $\eta$ & \ttt{eta} & 223 & $\omega$ & \ttt{omega} \\ 331 & $\eta'$ & \ttt{eta'} & 333 & $\phi$ & \ttt{phi} \\ 441 & $\eta_{\c}$ & \ttt{eta\_c} & 443 & $\mathrm{J}/\psi$ & \ttt{J/psi} \\ 551 & $\eta_{\b}$ & \ttt{eta\_b} & 553 & $\Upsilon$ & \ttt{Upsilon} \\ 661 & $\eta_{\t}$ & \ttt{eta\_t} & 663 & $\Theta$ & \ttt{Theta} \\ 130 & $\mathrm{K}_{\mrm{L}}^0$ & \ttt{K\_L0} & & & \\ 310 & $\mathrm{K}_{\mrm{S}}^0$ & \ttt{K\_S0} & & & \\ \hline \end{tabular} \end{center} \end{table} Also the lowest-lying orbital angular momentum $L = 1$ mesons are included, Table \ref{t:codesixb}: one pseudovector multiplet obtained for total quark-spin 0 ($L = 1, S = 0 \Rightarrow J = 1$) and one scalar, one pseudovector and one tensor multiplet obtained for total quark-spin 1 ($L = 1, S = 1 \Rightarrow J = 0, 1$ or 2), where $J$ is what is conventionally called the spin $s$ of the meson. Any mixing between the two pseudovector multiplets is not taken into account. Please note that some members of these multiplets have still not been found, and are included here only based on guesswork. Even for known ones, the information on particles (mass, width, decay modes) is highly incomplete. Only two radial excitations are included, the $\psi' = \psi(2S)$ and $\Upsilon' = \Upsilon(2S)$. \begin{table}[ptb] \captive{Meson codes, part 2. \protect\label{t:codesixb} } \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c||c|c|c|@{\protect\rule{0mm}{\tablinsep}}} \hline KF & Name & Printed & KF & Name & Printed \\ \hline 10213 & $\b_1$ & \ttt{b\_1+} & 10211 & $\a_0^+$ & \ttt{a\_0+} \\ 10313 & $\mathrm{K}_1^0$ & \ttt{K\_10} & 10311 & $\mathrm{K}_0^{*0}$ & \ttt{K*\_00} \\ 10323 & $\mathrm{K}_1^+$ & \ttt{K\_1+} & 10321 & $\mathrm{K}_0^{*+}$ & \ttt{K*\_0+} \\ 10413 & $\mathrm{D}_1^+$ & \ttt{D\_1+} & 10411 & $\mathrm{D}_0^{*+}$ & \ttt{D*\_0+} \\ 10423 & $\mathrm{D}_1^0$ & \ttt{D\_10} & 10421 & $\mathrm{D}_0^{*0}$ & \ttt{D*\_00} \\ 10433 & $\mathrm{D}_{1 \mathrm{s}}^+$ & \ttt{D\_1s+} & 10431 & $\mathrm{D}_{0 \mathrm{s}}^{*+}$ & \ttt{D*\_0s+} \\ 10113 & $\b_1^0$ & \ttt{b\_10} & 10111 & $\a_0^0$ & \ttt{a\_00} \\ 10223 & $\mathrm{h}_1^0$ & \ttt{h\_10} & 10221 & $\mathrm{f}_0^0$ & \ttt{f\_00} \\ 10333 & $\mathrm{h}'^0_1$ & \ttt{h'\_10} & 10331 & $\mathrm{f}'^0_0$ & \ttt{f'\_00} \\ 10443 & $\mathrm{h}_{1 \c}^0$ & \ttt{h\_1c0} & 10441 & $\chi_{0 \c}^0$ & \ttt{chi\_0c0} \\ \hline 20213 & $\a_1^+$ & \ttt{a\_1+} & 215 & $\a_2^+$ & \ttt{a\_2+} \\ 20313 & $\mathrm{K}_1^{*0}$ & \ttt{K*\_10} & 315 & $\mathrm{K}_2^{*0}$ & \ttt{K*\_20} \\ 20323 & $\mathrm{K}_1^{*+}$ & \ttt{K*\_1+} & 325 & $\mathrm{K}_2^{*+}$ & \ttt{K*\_2+} \\ 20413 & $\mathrm{D}_1^{*+}$ & \ttt{D*\_1+} & 415 & $\mathrm{D}_2^{*+}$ & \ttt{D*\_2+} \\ 20423 & $\mathrm{D}_1^{*0}$ & \ttt{D*\_10} & 425 & $\mathrm{D}_2^{*0}$ & \ttt{D*\_20} \\ 20433 & $\mathrm{D}_{1 \mathrm{s}}^{*+}$ & \ttt{D*\_1s+} & 435 & $\mathrm{D}_{2 \mathrm{s}}^{*+}$ & \ttt{D*\_2s+} \\ 20113 & $\a_1^0$ & \ttt{a\_10} & 115 & $\a_2^0$ & \ttt{a\_20} \\ 20223 & $\mathrm{f}_1^0$ & \ttt{f\_10} & 225 & $\mathrm{f}_2^0$ & \ttt{f\_20} \\ 20333 & $\mathrm{f}'^0_1$ & \ttt{f'\_10} & 335 & $\mathrm{f}'^0_2$ & \ttt{f'\_20} \\ 20443 & $\chi_{1 \c}^0$ & \ttt{chi\_1c0} & 445 & $\chi_{2 \c}^0$ & \ttt{chi\_2c0} \\ \hline 30443 & $\psi'$ & \ttt{psi'} & & & \\ 30553 & $\Upsilon'$ & \ttt{Upsilon'} & & & \\ \hline \end{tabular} \end{center} \end{table} The corresponding meson KC codes, used for organizing mass and decay data, range between 101 and 240. \item Baryon codes, Table \ref{t:codeseven}. \\ A baryon made up of quarks $i$, $j$ and $k$, with $i \geq j \geq k$, and total spin $s$, is given the code \begin{equation} \mrm{KF} = 1000 i + 100 j + 10 k + 2s + 1 ~. \end{equation} An exception is provided by spin $1/2$ baryons made up of three different types of quarks, where the two lightest quarks form a spin-0 diquark ($\Lambda$-like baryons). Here the order of the $j$ and $k$ quarks is reversed, so as to provide a simple means of distinction to baryons with the lightest quarks in a spin-1 diquark ($\Sigma$-like baryons). For hadrons with heavy flavours, the root names are Lambda or Sigma for hadrons with two $\u$ or $\d$ quarks, Xi for those with one, and Omega for those without $\u$ or $\d$ quarks. Some of the most frequently used codes are given in Table \ref{t:codeseven}. The full lowest-lying spin $1/2$ and $3/2$ multiplets are included in the program. The corresponding KC codes, used for organizing mass and decay data, range between 301 and 400, with some slots still free. \begin{table}[ptb] \captive{Baryon codes. \protect\label{t:codeseven} } \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c||c|c|c|@{\protect\rule{0mm}{\tablinsep}}} \hline KF & Name & Printed & KF & Name & Printed \\ \hline & & & 1114 & $\Delta^-$ & \ttt{Delta-} \\ 2112 & $\mathrm{n}$ & \ttt{n0} & 2114 & $\Delta^0$ & \ttt{Delta0} \\ 2212 & $\mathrm{p}$ & \ttt{p+} & 2214 & $\Delta^+$ & \ttt{Delta+} \\ & & & 2224 & $\Delta^{++}$ & \ttt{Delta++} \\ 3112 & $\Sigma^-$ & \ttt{Sigma-} & 3114 & $\Sigma^{*-}$ & \ttt{Sigma*-} \\ 3122 & $\Lambda^0$ & \ttt{Lambda0} & & & \\ 3212 & $\Sigma^0$ & \ttt{Sigma0} & 3214 & $\Sigma^{*0}$ & \ttt{Sigma*0} \\ 3222 & $\Sigma^+$ & \ttt{Sigma+} & 3224 & $\Sigma^{*+}$ & \ttt{Sigma*+} \\ 3312 & $\Xi^-$ & \ttt{Xi-} & 3314 & $\Xi^{*-}$ & \ttt{Xi*-} \\ 3322 & $\Xi^0$ & \ttt{Xi0} & 3324 & $\Xi^{*0}$ & \ttt{Xi*0} \\ & & & 3334 & $\Omega^-$ & \ttt{Omega-} \\ 4112 & $\Sigma_{\c}^0$ & \ttt{Sigma\_c0} & 4114 & $\Sigma_{\c}^{*0}$ & \ttt{Sigma*\_c0} \\ 4122 & $\Lambda_{\c}^+$ & \ttt{Lambda\_c+} & & & \\ 4212 & $\Sigma_{\c}^+$ & \ttt{Sigma\_c+} & 4214 & $\Sigma_{\c}^{*+}$ & \ttt{Sigma*\_c+} \\ 4222 & $\Sigma_{\c}^{++}$ & \ttt{Sigma\_c++} & 4224 & $\Sigma_{\c}^{*++}$ & \ttt{Sigma*\_c++} \\ 4132 & $\Xi_{\c}^0$ & \ttt{Xi\_c0} & & & \\ 4312 & $\Xi'^0_{\c}$ & \ttt{Xi'\_c0} & 4314 & $\Xi_{\c}^{*0}$ & \ttt{Xi*\_c0} \\ 4232 & $\Xi_{\c}^+$ & \ttt{Xi\_c+} & & & \\ 4322 & $\Xi'^+_{\c}$ & \ttt{Xi'\_c+} & 4324 & $\Xi_{\c}^{*+}$ & \ttt{Xi*\_c+} \\ 4332 & $\Omega_{\c}^0$ & \ttt{Omega\_c0} & 4334 & $\Omega_{\c}^{*0}$ & \ttt{Omega*\_c0} \\ 5112 & $\Sigma_{\b}^-$ & \ttt{Sigma\_b-} & 5114 & $\Sigma_{\b}^{*-}$ & \ttt{Sigma*\_b-} \\ 5122 & $\Lambda_{\b}^0$ & \ttt{Lambda\_b0} & & & \\ 5212 & $\Sigma_{\b}^0$ &\ttt{Sigma\_b0} & 5214 & $\Sigma_{\b}^{*0}$ & \ttt{Sigma*\_b0} \\ 5222 & $\Sigma_{\b}^+$ & \ttt{Sigma\_b+} & 5224 & $\Sigma_{\b}^{*+}$ & \ttt{Sigma*\_b+} \\ \hline \end{tabular} \end{center} \end{table} \item Diffractive states, Table \ref{t:codeeight}. \\ These codes are not standard ones: they have been defined by analogy to be used for denoting diffractive states in {\tsc{Pythia}}, as part of the event history. The first two or three digits give flavour content, while the last one is 0, to denote the somewhat unusual character of the code. Only a few codes have been introduced; depending on circumstances these also have to double up for other diffractive states. \begin{table}[ptb] \captive{Diffractive state codes. \protect\label{t:codeeight} } \\ \vspace{1ex} \begin{center} \begin{tabular}{|c|c|c|@{\protect\rule{0mm}{\tablinsep}}} \hline KF & Printed & Meaning \\ \hline 110 & \ttt{rho\_diff0} & Diffractive $\pi^0 / \rho^0 / \gamma$ state \\ 210 & \ttt{pi\_diffr+} & Diffractive $\pi^+$ state \\ 220 & \ttt{omega\_di0} & Diffractive $\omega$ state \\ 330 & \ttt{phi\_diff0} & Diffractive $\phi$ state \\ 440 & \ttt{J/psi\_di0} & Diffractive $\mathrm{J}/\psi$ state \\ 2110 & \ttt{n\_diffr} & Diffractive $\mathrm{n}$ state \\ 2210 & \ttt{p\_diffr+} & Diffractive $\mathrm{p}$ state \\ \hline \end{tabular} \end{center} \end{table} \item Free compressed codes. The positions 401--500 of mass and decay arrays are left open. Here a user may map any new kind of particle from the ordinary KF codes, which probably are above 10000, into a more manageable KC range for mass and decay data information. The mapping must be implemented in the \ttt{LUCOMP} function. \end{Enumerate} \subsection{The Event Record} \label{ss:evrec} Each new event generated is in its entirety stored in the common block \ttt{LUJETS}, which thus forms the event record. Here each jet or particle that appears at some stage of the fragmentation or decay chain will occupy one line in the matrices. The different components of this line will tell which jet/particle it is, from where it originates, its present status (fragmented/decayed or not), its momentum, energy and mass, and the space--time position of its production vertex. Note that \ttt{K(I,3)}--\ttt{K(I,5)} and the \ttt{P} and \ttt{V} vectors may take special meaning for some specific applications (e.g. sphericity or cluster analysis), as described in those connections. The event history information stored in \ttt{K(I,3)}--\ttt{K(I,5)} should not be taken too literally. In the particle decay chains, the meaning of a mother is well-defined, but the fragmentation description is more complicated. The primary hadrons produced in string fragmentation come from the string as a whole, rather than from an individual parton. Even when the string is not included in the history (see \ttt{MSTU(16)}), the pointer from hadron to parton is deceptive. For instance, in a $\mathrm{q}\mathrm{g}\overline{\mathrm{q}}$ event, those hadrons are pointing towards the $\mathrm{q}$ ($\overline{\mathrm{q}}$) parton that were produced by fragmentation from that end of the string, according to the random procedure used in the fragmentation routine. No particles point to the $\mathrm{g}$. This assignment seldom agrees with the visual impression, and is not intended to. The common block \ttt{LUJETS} has expanded with time, and can now house 4000 entries. This figure may seem ridiculously large, but actually the previous limit of 2000 was often reached in studies of high-$p_{\perp}$ processes at the LHC and SSC. This is because the event record contains not only the final particles, but also all intermediate partons and hadrons, which subsequenty showered, fragmented or decayed. Included are also a wealth of photons coming from $\pi^0$ decays; the simplest way of reducing the size of the event record is actually to switch off $\pi^0$ decays by \ttt{MDCY(LUCOMP(111),1)=0}. Also note that some routines, such as \ttt{LUCLUS} and \ttt{LUCELL}, use memory after the event record proper as a working area. Still, to change the size of the common block, upwards or downwards, is easy: just do a global substitute in the common block and change the \ttt{MSTU(4)} value to the new number. If more than 10000 lines are to be used, the packing of colour information should also be changed, see \ttt{MSTU(5)}. \drawbox{COMMON/LUJETS/N,K(4000,5),P(4000,5),V(4000,5)}\label{p:LUJETS} \begin{entry} \itemc{Purpose:} to contain the event record, i.e. the complete list of all partons and particles in the current event. \iteme{N :}\label{p:N} number of lines in the \ttt{K}, \ttt{P} and \ttt{V} matrices occupied by the current event. \ttt{N} is continuously updated as the definition of the original configuration and the treatment of fragmentation and decay proceed. In the following, the individual parton/particle number, running between 1 and \ttt{N}, is called \ttt{I}. \iteme{K(I,1) :}\label{p:K} status code KS, which gives the current status of the parton/particle stored in the line. The ground rule is that codes 1--10 correspond to currently existing partons/particles, while larger codes contain partons/particles which no longer exist, or other kinds of event information. \begin{subentry} \iteme{= 0 :} empty line. \iteme{= 1 :} an undecayed particle or an unfragmented jet, the latter being either a single jet or the last one of a jet system. \iteme{= 2 :} an unfragmented jet, which is followed by more jets in the same colour-singlet jet system. \iteme{= 3 :} an unfragmented jet with special colour flow information stored in \ttt{K(I,4)} and \ttt{K(I,5)}, such that adjacent partons along the string need not follow each other in the event record. \iteme{= 4 :} a particle which could have decayed, but did not within the allowed volume around the original vertex. \iteme{= 5 :} a particle which is to be forced to decay in the next \ttt{LUEXEC} call, in the vertex position given (this code is only set by user intervention). \iteme{= 11 :} a decayed particle or a fragmented jet, the latter being either a single jet or the last one of a jet system, cf. \ttt{=1}. \iteme{= 12 :} a fragmented jet, which is followed by more jets in the same colour-singlet jet system, cf. \ttt{=2}. Further, a $\mathrm{B}$ meson which decayed as a $\br{\mathrm{B}}$ one, or vice versa, because of $\mathrm{B}$--$\br{\mathrm{B}}$ mixing, is marked with this code rather than \ttt{=11}. \iteme{= 13 :} a jet which has been removed when special colour flow information has been used to rearrange a jet system, cf. \ttt{=3}. \iteme{= 14 :} a parton which has branched into further partons, with special colour-flow information provided, cf. \ttt{=3}. \iteme{= 15 :} a particle which has been forced to decay (by user intervention), cf. \ttt{=5}. \iteme{= 21 :} documentation lines used to give a compressed story of the event at the beginning of the event record. \iteme{= 31 :} lines with information on sphericity, thrust or cluster search. \iteme{= 32 :} tabular output, as generated by \ttt{LUTABU}. \iteme{= 41 :} junction (currently not fully implemented). \iteme{< 0 :} these codes are never used by the program, and are therefore usually not affected by operations on the record, such as \ttt{LUROBO}, \ttt{LULIST} and event-analysis routines (the exception is some \ttt{LUEDIT} calls, where lines are moved but not deleted). Such codes may therefore be useful in some connections. \end{subentry} \iteme{K(I,2) :} parton/particle KF code, as described in section \ref{ss:codes}. \iteme{K(I,3) :} line number of parent particle or jet, where known, otherwise 0. Note that the assignment of a particle to a given jet in a jet system is unphysical, and what is given there is only related to the way the event was generated. \iteme{K(I,4) :} normally the line number of the first daughter; it is 0 for an undecayed particle or unfragmented jet. For \ttt{K(I,1) = 3, 13} or \ttt{14}, instead, it contains special colour-flow information (for internal use only) of the form \\ \ttt{K(I,4)} = 200000000*MCFR + 100000000*MCTO + 10000*ICFR + ICTO, \\ where ICFR and ICTO give the line numbers of the partons from which the colour comes and to where it goes, respectively; MCFR and MCTO originally are 0 and are set to 1 when the corresponding colour connection has been traced in the \ttt{LUPREP} rearrangement procedure. (The packing may be changed with \ttt{MSTU(5)}.) The `from' colour position may indicate a parton which branched to produce the current parton, or a parton created together with the current parton but with matched anticolour, while the `to' normally indicates a parton that the current parton branches into. Thus, for setting up an initial colour configuration, it is normally only the `from' part that is used, while the `to' part is added by the program in a subsequent call to parton-shower evolution (for final-state radiation; it is the other way around for initial-state radiation). {\bf Note:} normally most users never have to worry about the exact rules for colour-flow storage, since this is used mainly for internal purposes. However, when it is necessary to define this flow, it is recommended to use the \ttt{LUJOIN} routine, since it is likely that this would reduce the chances of making a mistake. \iteme{K(I,5) :} normally the line number of the last daughter; it is 0 for an undecayed particle or unfragmented jet. For \ttt{K(I,1) = 3, 13} or \ttt{14}, instead, it contains special colour-flow information (for internal use only) of the form \\ \ttt{K(I,5)} = 200000000*MCFR + 100000000*MCTO + 10000*ICFR + ICTO, \\ where ICFR and ICTO give the line numbers of the partons from which the anticolour comes and to where it goes, respectively; MCFR and MCTO originally are 0 and are set to 1 when the corresponding colour connection has been traced in the \ttt{LUPREP} rearrangement procedure. For further discussion, see \ttt{K(I,4)}. \iteme{P(I,1) :}\label{p:P} $p_x$, momentum in the $x$ direction, in GeV/$c$. \iteme{P(I,2) :} $p_y$, momentum in the $y$ direction, in GeV/$c$. \iteme{P(I,3) :} $p_z$, momentum in the $z$ direction, in GeV/$c$. \iteme{P(I,4) :} $E$, energy, in GeV. \iteme{P(I,5) :} $m$, mass, in GeV/$c^2$. In parton showers, with space-like virtualities, i.e. where $Q^2 = - m^2 > 0$, one puts \ttt{P(I,5)}$ = -Q$. \iteme{V(I,1) :}\label{p:V} $x$ position of production vertex, in mm. \iteme{V(I,2) :} $y$ position of production vertex, in mm. \iteme{V(I,3) :} $z$ position of production vertex, in mm. \iteme{V(I,4) :} time of production, in mm/$c$ ($\approx 3.33 \times 10^{-12}$ s). \iteme{V(I,5) :} proper lifetime of particle, in mm/$c$ ($\approx 3.33 \times 10^{-12}$ s). If the particle is not expected to decay, \ttt{V(I,5)=0}. A line with \ttt{K(I,1)=4}, i.e. a particle that could have decayed, but did not within the allowed region, has the proper non-zero \ttt{V(I,5)}. In the absence of electric or magnetic fields, or other disturbances, the decay vertex \ttt{VP} of an unstable particle may be calculated as \\ \ttt{VP(j) = V(I,j) + V(I,5)*P(I,j)/P(I,5)}, \ttt{j} = 1--4. \end{entry} \subsection{How The Event Record Works} The event record is the main repository for information about an event. In the generation chain, it is used as a `scoreboard' for what has already been done and what remains to be done. This information can be studied by you, to access information not only about the final state, but also about what came before. \subsubsection{A simple example} The example of section \ref{ss:JETstarted} may help to clarify what is going on. When \ttt{LU2ENT} is called to generate a $\mathrm{q}\overline{\mathrm{q}}$ pair, the quarks are stored in lines 1 and 2 of the event record, respectively. Colour information is set to show that they belong together as a colour singlet. The counter \ttt{N} is also updated to the value of 2. At no stage is the previously generated event removed. Lines 1 and 2 are overwritten, but lines 3 onwards still contain whatever may have been there before. This does not matter, since \ttt{N} indicates where the `real' record ends. As \ttt{LUEXEC} is called, explicitly by you or indirectly by \ttt{LU2ENT}, the first entry is considered and found to be the first jet of a system. Therefore the second entry is also found, and these two together form a jet system, which may be allowed to fragment. The `string' that fragments is put in line 3 and the fragmentation products in lines 4 through 10 (in this particular case). At the same time, the $\mathrm{q}$ and $\overline{\mathrm{q}}$ in the first two lines are marked as having fragmented, and the same for the string. At this stage, \ttt{N} is 10. Internally there is another counter with the value 2, which indicates how far down in the record the event has been studied. This second counter is gradually increased by one. If the entry in the corresponding line can fragment or decay, then fragmentation or decay is perfomed. The fragmentation/decay products are added at the end of the event record, and \ttt{N} is updated accordingly. The entry is then also marked as having been treated. For instance, when line 3 is considered, the `string' entry of this line is seen to have been fragmented, and no action is taken. Line 4, a $\rho^+$, is allowed to decay to $\pi^+ \pi^0$; the decay products are stored in lines 11 and 12, and line 4 is marked as having decayed. Next, entry 5 is allowed to decay. The entry in line 6, $\pi^+$, is a stable particle (by default) and is therefore passed by without any action being taken. In the beginning of the process, entries are usually unstable, and \ttt{N} grows faster than the second counter of treated entries. Later on, an increasing fraction of the entries are stable end products, and the r\^oles are now reversed, with the second counter growing faster. When the two coincide, the end of the record has been reached, and the process can be stopped. All unstable objects have now been allowed to fragment or decay. They are still present in the record, so as to simplify the tracing of the history. Notice that \ttt{LUEXEC} could well be called a second time. The second counter would then start all over from the beginning, but slide through until the end without causing any action, since all objects that can be treated already have been. Unless some of the relevant switches were changed meanwhile, that is. For instance, if $\pi^0$ decays were switched off the first time around but on the second, all the $\pi^0$'s found in the record would be allowed to decay in the second call. A particle once decayed is not `undecayed', however, so if the $\pi^0$ is put back stable and \ttt{LUEXEC} is called a third time, nothing will happen. \subsubsection{Applications to PYTHIA} \label{sss:PYrecord} In a full-blown event generated with {\tsc{Pythia}}, the usage of \ttt{LUJETS} is more complicated, although the general principles survive. \ttt{LUJETS} is used extensively both by the {\tsc{Pythia}} and the {\tsc{Jetset}} routines; indeed it provides the bridge that allows the general utility routines in {\tsc{Jetset}} to be used also for {\tsc{Pythia}} events. The {\tsc{Pythia}} event listing begins (optionally) with a few lines of event summary, specific to the hard process simulated and thus not described in the overview above. These specific parts are covered in the following. In most instances, only the partons and particles actually produced are of interest. For \ttt{MSTP(125)=0}, the event record starts off with the parton configuration existing after hard interaction, initial- and final-state radiation, multiple interactions and beam remnants have been considered. The partons are arranged in colour singlet clusters, ordered as required for string fragmentation. Also photons and leptons produced as part of the hard interaction (e.g. from $\mathrm{q}\overline{\mathrm{q}} \to \mathrm{g} \gamma$ or $\u\overline{\mathrm{u}} \to \mathrm{Z}^0 \to \e^+\e^-$) appear in this part of the event record. These original entries appear with pointer \ttt{K(I,3)=0}, whereas the products of the subsequent fragmentation and decay have \ttt{K(I,3)} numbers pointing back to the line of the parent. The standard documentation, obtained with \ttt{MSTP(125)=1}, includes a few lines at the beginning of the event record, which contain a brief summary of the process that has taken place. The number of lines used depends on the nature of the hard process and is stored in \ttt{MSTI(4)} for the current event. These lines all have \ttt{K(I,1)=21}. For all processes, lines 1 and 2 give the two incoming hadrons. When listed with \ttt{LULIST}, these two lines will be separated from subsequent ones by a sequence of `\ttt{======}' signs, to improve readability. For diffractive and elastic events, the two outgoing states in lines 3 and 4 complete the list. Otherwise, lines 3 and 4 contain the two partons that initiate the two initial-state parton showers, and 5 and 6 the end products of these showers, i.e. the partons that enter the hard interaction. With initial-state radiation switched off, lines 3 and 5 and lines 4 and 6 coincide. For a simple $2 \to 2$ hard scattering, lines 7 and 8 give the two outgoing partons/particles from the hard interaction, before any final-state radiation. For $2 \to 2$ processes proceeding via an intermediate resonance such as $\gamma^* / \Z^0$, $\mathrm{W}^{\pm}$ or $\H^0$, the resonance is found in line 7 and the two outgoing partons/particles in 8 and 9. In some cases one of these may be a resonance in its own right, or both of them, so that further pairs of lines are added for subsequent decays. If the decay of a given resonance has been switched off, then no decay products are listed either in this initial summary or in the subsequent ordinary listing. Whenever partons are listed, they are assumed to be on the mass shell for simplicity. The fact that effective masses may be generated by initial- and final-state radiation is taken into account in the actual parton configuration that is allowed to fragment, however. A special case is provided by $\mathrm{W}^+ \mathrm{W}^-$ or $\mathrm{Z}^0 \mathrm{Z}^0$ fusion to an $\H^0$. Then the virtual $\mathrm{W}$'s or $\mathrm{Z}$'s are shown in lines 7 and 8, the $\H^0$ in line 9, and the two recoiling quarks (that emitted the bosons) in 10 and 11, followed by the Higgs decay products. Since the $\mathrm{W}$'s and $\mathrm{Z}$'s are space-like, what is actually listed as the mass for them is $-\sqrt{-m^2}$. The listing of the event documentation closes with another line made up of `\ttt{======}' signs. A few examples may help clarify the picture. For a single diffractive event $\mathrm{p} \overline{\mathrm{p}} \to \mathrm{p}_{\mrm{diffr}} \overline{\mathrm{p}}$, the event record will start with \\ \verb& I K(I,1) K(I,2) K(I,3) & comment \\ \verb& 1 21 2212 0 & incoming $\mathrm{p}$ \\ \verb& 2 21 -2212 0 & incoming $\overline{\mathrm{p}}$ \\ \verb&========================= & not part of record; appears in listings \\ \verb& 3 21 27 1 & outgoing $\mathrm{p}_{\mrm{diffr}}$ \\ \verb& 4 21 -2212 2 & outgoing $\overline{\mathrm{p}}$ \\ \verb&========================= & again not part of record The typical QCD $2 \to 2$ process would be \\ \verb& I K(I,1) K(I,2) K(I,3) & comment \\ \verb& 1 21 2212 0 & incoming $\mathrm{p}$ \\ \verb& 2 21 -2212 0 & incoming $\overline{\mathrm{p}}$ \\ \verb&========================= & \\ \verb& 3 21 2 1 & $\u$ picked from incoming $\mathrm{p}$ \\ \verb& 4 21 -1 2 & $\overline{\mathrm{d}}$ picked from incoming $\overline{\mathrm{p}}$ \\ \verb& 5 21 21 3 & $\u$ evolved to $\mathrm{g}$ at hard scattering \\ \verb& 6 21 -1 4 & still $\overline{\mathrm{d}}$ at hard scattering \\ \verb& 7 21 21 0 & outgoing $\mathrm{g}$ from hard scattering \\ \verb& 8 21 -1 0 & outgoing $\overline{\mathrm{d}}$ from hard scattering \\ \verb&========================= & Note that, where well defined, the \ttt{K(I,3)} code does contain information as to which side the different partons come from, e.g. above the gluon in line 5 points back to the $\u$ in line 3, which points back to the proton in line 1. In the example above, it would have been possible to associate the scattered g in line 7 with the incoming one in line 5, but this is not possible in the general case, consider e.g. $\mathrm{g} \mathrm{g} \to \mathrm{g} \mathrm{g}$. As a final example, $\mathrm{W}^+\mathrm{W}^-$ fusion to an $\H^0$ in process 8 (not process 124, which is lengthier) might look like \\ \verb& I K(I,1) K(I,2) K(I,3) & comment \\ \verb& 1 21 2212 0 & first incoming $\mathrm{p}$ \\ \verb& 2 21 2212 0 & second incoming $\mathrm{p}$ \\ \verb&========================= & \\ \verb& 3 21 2 1 & $\u$ picked from first $\mathrm{p}$ \\ \verb& 4 21 21 2 & $\mathrm{g}$ picked from second $\mathrm{p}$ \\ \verb& 5 21 2 3 & still $\u$ after initial-state radiation \\ \verb& 6 21 -4 4 & $\mathrm{g}$ evolved to $\overline{\mathrm{c}}$ \\ \verb& 7 21 24 5 & space-like $\mathrm{W}^+$ emitted by $\u$ quark \\ \verb& 8 21 -24 6 & space-like $\mathrm{W}^-$ emitted by $\overline{\mathrm{c}}$ quark \\ \verb& 9 21 25 0 & Higgs produced by $\mathrm{W}^+ \mathrm{W}^-$ fusion \\ \verb&10 21 1 5 & $\u$ turned into $\d$ by emission of $\mathrm{W}^+$ \\ \verb&11 21 -3 6 & $\overline{\mathrm{c}}$ turned into $\overline{\mathrm{s}}$ by emission of $\mathrm{W}^-$ \\ \verb&12 21 23 9 & first $\mathrm{Z}^0$ coming from decay of $\H^0$ \\ \verb&13 21 23 9 & second $\mathrm{Z}^0$ coming from decay of $\H^0$ \\ \verb&14 21 12 12 & $\nu_{\mathrm{e}}$ from first $\mathrm{Z}^0$ decay \\ \verb&15 21 -12 12 & $\br{\nu}_{\mathrm{e}}$ from first $\mathrm{Z}^0$ decay \\ \verb&16 21 5 13 & $\b$ quark from second $\mathrm{Z}^0$ decay \\ \verb&17 21 -5 13 & $\overline{\mathrm{b}}$ antiquark from second $\mathrm{Z}^0$ decay \\ \verb&========================= & After these lines with the initial information, the event record looks the same as for \ttt{MSTP(125)=0}, i.e. first comes the parton configuration to be fragmented and, after another separator line `\ttt{======}' in the output (but not the event record), the products of subsequent fragmentation and decay chains. The \ttt{K(I,3)} pointers for the partons, as well as leptons and photons produced in the hard interaction, are now pointing towards the documentation lines above, however. In particular, beam remnants point to 1 or 2, depending on which side they belong to, and partons emitted in the initial-state parton showers point to 3 or 4. In the second example above, the partons produced by final-state radiation will be pointing back to 7 and 8; as usual, it should be remembered that a specific assignment to 7 or 8 need not be unique. For the third example, final-state radiation partons will come both from partons 10 and 11 and from partons 16 and 17, and additionally there will be a neutrino--antineutrino pair pointing to 14 and 15. The extra pairs of partons that are generated by multiple interactions do not point back to anything, i.e. they have \ttt{K(I,3)=0}. There exists a third documentation option, \ttt{MSTP(125)=2}. Here the history of initial- and final-state parton branchings may be traced, including all details on colour flow. This information has not been optimized for user-friendliness, and cannot be recommended for general usage. With this option, the initial documentation lines are the same. They are followed by blank lines, \ttt{K(I,1)=0}, up to line 20 (can be changed in \ttt{MSTP(126)}). From line 21 onwards each parton with \ttt{K(I,1)=} 3, 13 or 14 appears with special colour-flow information in the \ttt{K(I,4)} and \ttt{K(I,5)} positions. For an ordinary $2 \to 2$ scattering, the two incoming partons at the hard scattering are stored in lines 21 and 22, and the two outgoing in 23 and 24. The colour flow between these partons has to be chosen according to the proper relative probabilities in cases when many alternatives are possible, see section \ref{sss:QCDjetclass}. If there is initial-state radiation, the two partons in lines 21 and 22 are copied down to lines 25 and 26, from which the initial-state showers are reconstructed backwards step by step. The branching history may be read by noting that, for a branching $a \to b c$, the \ttt{K(I,3)} codes of $b$ and $c$ point towards the line number of $a$. Since the showers are reconstructed backwards, this actually means that parton $b$ would appear in the listing before parton $a$ and $c$, and hence have a pointer to a position below itself in the list. Associated time-like partons $c$ may initiate time-like showers, as may the partons of the hard scattering. Again a showering parton or pair of partons will be copied down towards the end of the list and allowed to undergo successive branchings $c \to d e$, with $d$ and $e$ pointing towards $c$. The mass of time-like partons is properly stored in \ttt{P(I,5)}; for space-like partons $-\sqrt{-m^2}$ is stored instead. After this section, containing all the branchings, comes the final parton configuration, properly arranged in colour, followed by all subsequent fragmentation and decay products, as usual. \subsection{The HEPEVT Standard} \label{ss:HEPEVT} A set of common blocks was developed and agreed on within the framework of the 1989 LEP physics study, see \cite{Sjo89}. This standard defines an event record structure which should make the interfacing of different event generators much simpler. It would be a major work to rewrite {\tsc{Pythia/Jetset}} to agree with this standard event record structure. More importantly, the standard only covers quantities which can be defined unambiguously, i.e. which are independent of the particular program used. There are thus no provisions for the need for colour-flow information in models based on string fragmentation, etc., so the standard common blocks would anyway have to be supplemented with additional event information. For the moment, the adopted approach is therefore to retain the \ttt{LUJETS} event record, but supply a routine \ttt{LUHEPC} which can convert to or from the standard event record. Owing to a somewhat different content in the two records, some ambiguities do exist in the translation procedure. \ttt{LUHEPC} has therefore to be used with some judgment. In this section, the new standard event structure is first presented, i.e. the most important points in \cite{Sjo89} are recapitulated. Thereafter the conversion routine is described, with particular attention to ambiguities and limitations. The standard event record is stored in two common blocks. The second of these is specifically intended for spin information. Since {\tsc{Jetset}} never (explicitly) makes use of spin information, this latter common block is not addressed here. A third common block for colour flow information has been discussed, but never formalized. In order to make the components of the standard more distinguishable in user programs, the three characters \ttt{HEP} (for High Energy Physics) have been chosen to be a part of all names. Originally it was not specified whether real variables should be in single or double precision. At the time, this meant that single precision became the default choice, but since then the trend has been towards increasing precision. In connection with the 1995 LEP~2 workshop, it was therefore agreed to adopt \ttt{DOUBLE PRECISION} real variables as part of the standard. \drawboxfour{~PARAMETER (NMXHEP=2000)} {~COMMON/HEPEVT/NEVHEP,NHEP,ISTHEP(NMXHEP),IDHEP(NMXHEP),} {\&JMOHEP(2,NMXHEP),JDAHEP(2,NMXHEP),PHEP(5,NMXHEP),VHEP(4,NMXHEP)} {~DOUBLE PRECISION PHEP, VHEP} \label{p:HEPEVT}\begin{entry} \itemc{Purpose:} to contain an event record in a Monte Carlo-independent format. \iteme{NMXHEP:} maximum numbers of entries (partons/particles) that can be stored in the common block. The default value of 2000 can be changed via the parameter construction. In the translation, it is checked that this value is not exceeded. \iteme{NEVHEP:} is normally the event number, but may have special meanings, according to the description below: \begin{subentry} \iteme{> 0 :} event number, sequentially increased by 1 for each call to the main event generation routine, starting with 1 for the first event generated. \iteme{= 0 :} for a program which does not keep track of event numbers, as {\tsc{Jetset}}. \iteme{= -1 :} special initialization record; not used by {\tsc{Jetset}}. \iteme{= -2 :} special final record; not used by {\tsc{Jetset}}. \end{subentry} \iteme{NHEP:} the actual number of entries stored in the current event. These are found in the first \ttt{NHEP} positions of the respective arrays below. Index \ttt{IHEP}, 1$\leq$\ttt{IHEP}$\leq$\ttt{NHEP}, is used below to denote a given entry. \iteme{ISTHEP(IHEP):} status code for entry \ttt{IHEP}, with the following meanings: \begin{subentry} \iteme{= 0 :} null entry. \iteme{= 1 :} an existing entry, which has not decayed or fragmented. This is the main class of entries, which represents the `final state' given by the generator. \iteme{= 2 :} an entry which has decayed or fragmented and is therefore not appearing in the final state, but is retained for event history information. \iteme{= 3 :} a documentation line, defined separately from the event history. This could include the two incoming reacting particles, etc. \iteme{= 4 - 10 :} undefined, but reserved for future standards. \iteme{= 11 - 200 :} at the disposal of each model builder for constructs specific to his program, but equivalent to a null line in the context of any other program. \iteme{= 201 - :} at the disposal of users, in particular for event tracking in the detector. \end{subentry} \iteme{IDHEP(IHEP) :} particle identity, according to the PDG standard. The four additional codes 91--94 have been introduced to make the event history more legible, see section \ref{ss:codes} and the \ttt{MSTU(16)} description. \iteme{JMOHEP(1,IHEP) :} pointer to the position where the mother is stored. The value is 0 for initial entries. \iteme{JMOHEP(2,IHEP) :} pointer to position of second mother. Normally only one mother exists, in which case the value 0 is to be used. In {\tsc{Jetset}}, entries with codes 91--94 are the only ones to have two mothers. The flavour contents of these objects, as well as details of momentum sharing, have to be found by looking at the mother partons, i.e. the two partons in positions \ttt{JMOHEP(1,IHEP)} and \ttt{JMOHEP(2,IHEP)} for a cluster or a shower system, and the range \ttt{JMOHEP(1,IHEP)}--\ttt{JMOHEP(2,IHEP)} for a string or an independent fragmentation parton system. \iteme{JDAHEP(1,IHEP) :} pointer to the position of the first daughter. If an entry has not decayed, this is 0. \iteme{JDAHEP(2,IHEP) :} pointer to the position of the last daughter. If an entry has not decayed, this is 0. It is assumed that daughters are stored sequentially, so that the whole range \ttt{JDAHEP(1,IHEP)}--\ttt{JDAHEP(2,IHEP)} contains daughters. This variable should be set also when only one daughter is present, as in $\mathrm{K}^0 \to \mathrm{K}_{\mrm{S}}^0$ decays, so that looping from the first daughter to the last one works transparently. Normally daughters are stored after mothers, but in backwards evolution of initial-state radiation the opposite may appear, i.e. that mothers are found below the daughters they branch into. Also, the two daughters then need not appear one after the other, but may be separated in the event record. \iteme{PHEP(1,IHEP) :} momentum in the $x$ direction, in GeV/$c$. \iteme{PHEP(2,IHEP) :} momentum in the $y$ direction, in GeV/$c$. \iteme{PHEP(3,IHEP) :} momentum in the $z$ direction, in GeV/$c$. \iteme{PHEP(4,IHEP) :} energy, in GeV. \iteme{PHEP(5,IHEP) :} mass, in GeV/$c^2$. For space-like partons, it is allowed to use a negative mass, according to \ttt{PHEP(5,IHEP)}$ = -\sqrt{-m^2}$. \iteme{VHEP(1,IHEP) :} production vertex $x$ position, in mm. \iteme{VHEP(2,IHEP) :} production vertex $y$ position, in mm. \iteme{VHEP(3,IHEP) :} production vertex $z$ position, in mm. \iteme{VHEP(4,IHEP) :} production time, in mm/$c$ ($\approx 3.33 \times 10^{-12}$ s). \end{entry} \boxsep This completes the brief description of the standard. In {\tsc{Jetset}}, the routine \ttt{LUHEPC} is provided as an interface. \drawbox{CALL LUHEPC(MCONV)}\label{p:LUHEPC} \begin{entry} \itemc{Purpose:} to convert between the \ttt{LUJETS} event record and the \ttt{HEPEVT} event record. \iteme{MCONV :} direction of conversion. \begin{subentry} \iteme{= 1 :} translates the current \ttt{LUJETS} record into the \ttt{HEPEVT} one, while leaving the original \ttt{LUJETS} one unaffected. \iteme{= 2 :} translates the current \ttt{HEPEVT} record into the \ttt{LUJETS} one, while leaving the original \ttt{HEPEVT} one unaffected. \end{subentry} \end{entry} \boxsep The conversion of momenta is trivial: it is just a matter of exchanging the order of the indices. The vertex information is but little more complicated; the extra fifth component present in \ttt{LUJETS} can be easily reconstructed from other information for particles which have decayed. (Some of the advanced features made possible by this component, such as the possibility to consider decays within expanding spatial volumes in subsequent \ttt{LUEXEC} calls, cannot be used if the record is translated back and forth, however.) Also, the particle codes \ttt{K(I,2)} and \ttt{IDHEP(I)} are identical, since they are both based on the PDG codes. The remaining, non-trivial areas deal with the status codes and the event history. In moving from \ttt{LUJETS} to \ttt{HEPEVT}, information on colour flow is lost. On the other hand, the position of a second mother, if any, has to be found; this only affects lines with \ttt{K(I,2)=} 91--94. Also, for lines with \ttt{K(I,1)=} 13 or 14, the daughter pointers have to be found. By and large, however, the translation from \ttt{LUJETS} to \ttt{HEPEVT} should cause little problem, and there should never be any need for user intervention. (We assume that {\tsc{Jetset}} is run with the default \ttt{MSTU(16)=1}, otherwise some discrepancies with respect to the proposed standard event history description will be present.) In moving from \ttt{HEPEVT} to \ttt{LUJETS}, information on a second mother is lost. Any codes \ttt{IDHEP(I)} not equal to 1, 2 or 3 are translated into \ttt{K(I,1)=0}, and so all entries with \ttt{K(I,1)}$\geq 30$ are effectively lost in a translation back and forth. All entries with \ttt{IDHEP(I)=2} are translated into \ttt{K(I,1)=11}, and so entries of type \ttt{K(I,1) = 12, 13, 14} or \ttt{15} are never found. There is thus no colour-flow information available for partons which have fragmented. For partons with \ttt{IDHEP(I)=1}, i.e. which have not fragmented, an attempt is made to subdivide the partonic system into colour singlets, as required for subsequent string fragmentation. To this end, it is assumed that partons are stored sequentially along strings. Normally, a string would then start at a $\mathrm{q}$ ($\overline{\mathrm{q}}$) or $\overline{\mathrm{q}}\qbar$ ($\mathrm{q}\q$) entry, cover a number of intermediate gluons, and end at a $\overline{\mathrm{q}}$ ($\mathrm{q}$) or $\mathrm{q}\q$ ($\overline{\mathrm{q}}\qbar$) entry. Particles could be interspersed in this list with no adverse effects, i.e. a $\u-\mathrm{g}-\gamma-\overline{\mathrm{u}}$ sequence would be interpreted as a $\u-\mathrm{g}-\overline{\mathrm{u}}$ string plus an additional photon. A closed gluon loop would be assumed to be made up of a sequential listing of the gluons, with the string continuing from the last gluon up back to the first one. Contrary to the previous, open string case, the appearance of any particle but a gluon would therefore signal the end of the gluon loop. For example, a $\mathrm{g}-\mathrm{g}-\mathrm{g}-\mathrm{g}$ sequence would be interpreted as one single four-gluon loop, while a $\mathrm{g}-\mathrm{g}-\gamma-\mathrm{g}-\mathrm{g}$ sequence would be seen as composed of two 2-gluon systems. If these interpretations, which are not unique, are not to your liking, it is up to you to correct them, e.g. by using \ttt{LUJOIN} to tell exactly which partons should be joined, in which sequence, to give a string. Calls to \ttt{LUJOIN} (or the equivalent) are also necessary if \ttt{LUSHOW} is to be used to have some partons develop a shower. For practical applications, one should note that {\tsc{Jetset}} $\e^+\e^-$ events, which have been allowed to shower but not to fragment, do have partons arranged in the order assumed above, so that a translation to \ttt{HEPEVT} and back does not destroy the possibility to perform fragmentation by a simple \ttt{LUEXEC} call. Also the hard interactions in {\tsc{Pythia}} fulfil this condition, while problems may appear in the multiple interaction scenario, where several closed $\mathrm{g}\g$ loops may appear directly following one another, and thus would be interpreted as a single multigluon loop after translation back and forth. \clearpage \section{Hard Processes in JETSET} \label{s:JETSETproc} {\tsc{Jetset}} contains the simulation of two hard processes. The process of main interest is $\e^+\e^- \to \gamma^* / \Z^0 \to \mathrm{q} \overline{\mathrm{q}}$. Higher-order QCD corrections can be obtained either with parton showers or with second-order matrix elements. The details of the parton-shower evolution are given in section \ref{s:showinfi}, while this section contains the matrix-element description, including a summary of the {\tsc{Jetset}} algorithm for initial-state photon radiation. Also {\tsc{Pythia}} can be used to simulate the process $\e^+\e^- \to \gamma^* / \Z^0 \to \mathrm{q} \overline{\mathrm{q}}$, but without the options of using second-order matrix elements or polarized incoming beams. Some other differences between the two algorithms are described. The other hard process in {\tsc{Jetset}} is $\Upsilon$ decay to $\mathrm{g} \mathrm{g} \mathrm{g}$ or $\gamma \mathrm{g} \mathrm{g}$, which is briefly commented on. The main sources of information for this chapter are refs. \cite{Sjo83,Sjo86,Sjo89}. \subsection{Annihilation Events in the Continuum} \label{ss:eematrix} The description of $\e^+\e^-$ annihilation into hadronic events involves a number of components: the $s$ dependence of the total cross section and flavour composition, multijet matrix elements, angular orientation of events, initial-state photon bremsstrahlung and effects of initial-state electron polarization. Many of the published formulae have been derived for the case of massless outgoing quarks. For each of the components described in the following, we will begin by discussing the massless case, and then comment on what is done to accommodate massive quarks. \subsubsection{Electroweak cross sections} In the standard theory, fermions have the following couplings (illustrated here for the first generation): \begin{center} \begin{tabular}{lll@{\protect\rule{0mm}{\tablinsep}}} $e_{\nu} = 0$, & $v_{\nu} = 1$, & $a_{\nu} = 1$, \\ $e_{\mathrm{e}} = -1$, & $v_{\mathrm{e}} = -1 + 4\sin^2 \! \theta_W$, & $a_{\mathrm{e}} = -1$, \\ $e_{\u} = 2/3$, & $v_{\u} = 1 - 8\sin^2 \! \theta_W /3$, & $a_{\nu} = 1$, \\ $e_{\d} = -1/3$, & $v_{\d} = -1 + 4\sin^2 \! \theta_W /3$, & $a_{\d} = -1$, \\ \end{tabular} \end{center} with $e$ the electric charge, and $v$ and $a$ the vector and axial couplings to the $\mathrm{Z}^0$. The relative energy dependence of the weak neutral current to the electromagnetic one is given by \begin{equation} \chi(s) = \frac{1}{4\sin^2 \! \theta_W\cos^2 \! \theta_W} \; \frac{s}{s - m_{\mathrm{Z}}^2 + i m_{\mathrm{Z}}\Gamma_{\mathrm{Z}}} ~, \label{ee:chis} \end{equation} where $s = E_{\mrm{cm}}^2$. In {\tsc{Jetset}} the electroweak mixing parameter $\sin^2 \! \theta_W$ and the $\mathrm{Z}^0$ mass $m_{\mathrm{Z}}$ and width $\Gamma_{\mathrm{Z}}$ are considered as constants to be given by you (while {\tsc{Pythia}} itself calculates an $s$-dependent width). Although the incoming $\mathrm{e}^+$ and $\mathrm{e}^-$ beams are normally unpolarized, we have included the possibility of polarized beams, following the formalism of \cite{Ols80}. Thus the incoming $\mathrm{e}^+$ and $\mathrm{e}^-$ are characterized by polarizations $\mbf{P}^{\pm}$ in the rest frame of the particles: \begin{equation} \mbf{P}^{\pm} = P_{\mrm{T}}^{\pm} \hat{\mbf{s}}^{\pm} + P_{\mrm{L}}^{\pm} \hat{\mbf{p}}^{\pm} ~, \end{equation} where $0 \leq P_{\mrm{T}}^{\pm} \leq 1$ and $-1 \leq P_{\mrm{L}}^{\pm} \leq 1$, with the constraint \begin{equation} (\mbf{P}^{\pm})^2 = (P_{\mrm{T}}^{\pm})^2 + (P_{\mrm{L}}^{\pm})^2 \leq 1 ~. \end{equation} Here $\hat{\mbf{s}}^{\pm}$ are unit vectors perpendicular to the beam directions $\hat{\mbf{p}}^{\pm}$. To be specific, we choose a right-handed coordinate frame with $\hat{\mbf{p}}^{\pm} = (0,0, \mp 1)$, and standard transverse polarization directions (out of the machine plane for storage rings) $\hat{\mbf{s}}^{\pm} = (0, \pm 1,0)$, the latter corresponding to azimuthal angles $\varphi^{\pm} = \pm \pi /2$. As free parameters in the program we choose $P_{\mrm{L}}^+$, $P_{\mrm{L}}^-$, $P_{\mrm{T}} = \sqrt{P_{\mrm{T}}^+ P_{\mrm{T}}^-}$ and $\Delta \varphi = (\varphi^+ + \varphi^-) /2$. In the massless QED case, the probability to produce a flavour $\mathrm{f}$ is proportional to $e_{\mathrm{f}}^2$, i.e up-type quarks are four times as likely as down-type ones. In lowest-order massless QFD the corresponding relative probabilities are given by \cite{Ols80} \begin{eqnarray} h_{\mathrm{f}}(s) & = & e_{\mathrm{e}}^2 \, (1 - P_{\mrm{L}}^+ P_{\mrm{L}}^-) \, e_{\mathrm{f}}^2 \, + \, 2 e_{\mathrm{e}} \left\{ v_{\mathrm{e}} (1 - P_{\mrm{L}}^+ P_{\mrm{L}}^-) - a_{\mathrm{e}} (P_{\mrm{L}}^- - P_{\mrm{L}}^+) \right\} \, \Re\chi(s) \, e_{\mathrm{f}} v_{\mathrm{f}} \, + \nonumber \\ & & + \, \left\{ (v_{\mathrm{e}}^2 + a_{\mathrm{e}}^2) (1 - P_{\mrm{L}}^+ P_{\mrm{L}}^-) - 2 v_{\mathrm{e}} a_{\mathrm{e}} (P_{\mrm{L}}^- - P_{\mrm{L}}^+) \right\} \, \left| \chi(s) \right|^2 \, \left\{ v_{\mathrm{f}}^2 + a_{\mathrm{f}}^2 \right\} ~, \label{ee:hf} \end{eqnarray} where $\Re\chi(s)$ denotes the real part of $\chi(s)$. The $h_{\mathrm{f}}(s)$ expression depends both on the $s$ value and on the longitudinal polarization of the $\mathrm{e}^{\pm}$ beams in a non-trivial way. The cross section for the process $\e^+\e^- \to \gamma^* / \Z^0 \to \mathrm{f} \overline{\mathrm{f}}$ may now be written as \begin{equation} \sigma_{\mathrm{f}}(s) = \frac{4 \pi \alpha_{\mathrm{em}}^2}{3 s} R_{\mathrm{f}}(s) ~, \end{equation} where $R_{\mathrm{f}}$ gives the ratio to the lowest-order QED cross section for the process $\e^+\e^- \to \mu^+ \mu^-$, \begin{equation} R_{\mathrm{f}}(s) = N_C \, R_{\mrm{QCD}} \, h_{\mathrm{f}}(s) ~. \end{equation} The factor of $N_C = 3$ counts the number of colour states available for the $\mathrm{q}\overline{\mathrm{q}}$ pair. The $R_{\mrm{QCD}}$ factor takes into account QCD loop corrections to the cross section. For $n_f$ effective flavours (normally $n_f =5$) \begin{equation} R_{\mrm{QCD}} \approx 1 + \frac{\alpha_{\mathrm{s}}}{\pi} + (1.986 - 0.115 n_f) \left( \frac{\alpha_{\mathrm{s}}}{\pi} \right)^2 + \cdots \label{ee:RQCD} \end{equation} in the $\br{\mrm{MS}}$ renormalization scheme \cite{Din79}. Note that $R_{\mrm{QCD}}$ does not affect the relative quark-flavour composition, and so is of peripheral interest in {\tsc{Jetset}}. (For leptons the $N_C$ and $R_{\mrm{QCD}}$ factors would be absent, i.e. $N_C \, R_{\mrm{QCD}} = 1$, but leptonic final states are not generated in {\tsc{Jetset}}.) Neglecting higher-order QCD and QFD effects, the corrections for massive quarks are given in terms of the velocity $v_{\mathrm{q}}$ of a quark with mass $m_{\mathrm{q}}$, $v_{\mathrm{q}} = \sqrt{ 1 - 4 m_{\mathrm{q}}^2 /s}$, as follows. The vector quark current terms in $h_{\mathrm{f}}$ (proportional to $e_{\mathrm{f}}^2$, $e_{\mathrm{f}} v_{\mathrm{f}}$, or $v_{\mathrm{f}}^2$) are multiplied by a threshold factor $v_{\mathrm{q}} (3 - v_{\mathrm{q}}^2) /2$, while the axial vector quark current term (proportional to $a_{\mathrm{f}}^2$) is multiplied by $v_{\mathrm{q}}^3$. While inclusion of quark masses in the QFD formulae decreases the total cross section, first-order QCD corrections tend in the opposite direction \cite{Jer81}. Na\"{\i}vely, one would expect one factor of $v_{\mathrm{q}}$ to get cancelled. So far, the available options are either to include threshold factors in full or not at all. Given that all five quarks are light at the scale of the $\mathrm{Z}^0$, the issue of quark masses is not really of interest at LEP. Here, however, purely weak corrections are important, in particular since they change the $\b$ quark partial width differently from that of the other ones \cite{Kuh89}. No such effects are included in the program. \subsubsection{First-order QCD matrix elements} The Born process $\e^+\e^- \to \mathrm{q} \overline{\mathrm{q}}$ is modified in first-order QCD by the probability for the $\mathrm{q}$ or $\overline{\mathrm{q}}$ to radiate a gluon, i.e. by the process $\e^+\e^- \to \mathrm{q} \overline{\mathrm{q}} \mathrm{g}$. The matrix element is conveniently given in terms of scaled energy variables in the c.m. frame of the event, $x_1 = 2E_{\mathrm{q}}/E_{\mrm{cm}}$, $x_2 = 2E_{\overline{\mathrm{q}}}/E_{\mrm{cm}}$, and $x_3 = 2E_{\mathrm{g}}/E_{\mrm{cm}}$, i.e. $x_1 + x_2 + x_3 = 2$. For massless quarks the matrix element reads \cite{Ell76} \begin{equation} \frac{1}{\sigma_0} \, \frac{\d \sigma}{\d x_1 \, \d x_2} = \frac{\alpha_{\mathrm{s}}}{2\pi} \, C_F \, \frac{x_1^2 + x_2^2}{(1-x_1)(1-x_2)} ~, \label{ee:ME3j} \end{equation} where $\sigma_0$ is the lowest-order cross section, $C_F = 4/3$ is the appropriate colour factor, and the kinematically allowed region is $0 \leq x_i \leq 1, i = 1, 2, 3$. By kinematics, the $x_k$ variable for parton $k$ is related to the invariant mass $m_{ij}$ of the other two partons $i$ and $j$ by $y_{ij} = m_{ij}^2/E_{\mrm{cm}}^2 = 1 - x_k$. The strong coupling constant $\alpha_{\mathrm{s}}$ is in first order given by \begin{equation} \alpha_{\mathrm{s}}(Q^2) = \frac{12\pi}{(33-2n_f) \, \ln(Q^2/\Lambda^2)} ~. \label{ee:aS3j} \end{equation} Conventionally $Q^2 = s = E_{\mrm{cm}}^2$; we will return to this issue below. The number of flavours $n_f$ is 5 for LEP applications, and so the $\Lambda$ value determined is $\Lambda_5$ (while e.g. most deep inelastic scattering studies refer to $\Lambda_4$, the energies for these experiments being below the bottom threshold). The $\alpha_{\mathrm{s}}$ values are matched at flavour thresholds, i.e. as $n_f$ is changed the $\Lambda$ value is also changed. It is therefore the derivative of $\alpha_{\mathrm{s}}$ that changes at a threshold, not $\alpha_{\mathrm{s}}$ itself. In order to separate 2-jets from 3-jets, it is useful to introduce jet-resolution parameters. This can be done in several different ways. Most famous are the $y$ and $(\epsilon, \delta)$ procedures. We will only refer to the $y$ cut, which is the one used in the program. Here a 3-parton configuration is called a 2-jet event if \begin{equation} \min_{i,j} (y_{ij}) = \min_{i,j} \left( \frac{m_{ij}^2}{E_{\mrm{cm}}^2} \right) < y ~. \end{equation} The cross section in eq.~(\ref{ee:ME3j}) diverges for $x_1 \rightarrow 1$ or $x_2 \rightarrow 1$ but, when first-order propagator and vertex corrections are included, a corresponding singularity with opposite sign appears in the $\mathrm{q} \overline{\mathrm{q}}$ cross section, so that the total cross section is finite. In analytical calculations, the average value of any well-behaved quantity ${\cal Q}$ can therefore be calculated as \begin{equation} \left\langle {\cal Q} \right\rangle = \frac{1}{\sigma_{\mrm{tot}}} \lim_{y \rightarrow 0} \left( {\cal Q}(\mrm{2parton}) \, \sigma_{\mrm{2parton}}(y) + \int_{y_{ij} > y} {\cal Q}(x_1,x_2) \, \frac{\d \sigma_{\mrm{3parton}}}{\d x_1 \, \d x_2} \, \d x_1 \, \d x_2 \right) ~, \label{ee:Obs} \end{equation} where any explicit $y$ dependence disappears in the limit $y \rightarrow 0$. In a Monte Carlo program, it is not possible to work with a negative total 2-jet rate, and thus it is necessary to introduce a fixed non-vanishing $y$ cut in the 3-jet phase space. Experimentally, there is evidence for the need of a low $y$ cut, i.e. a large 3-jet rate. For LEP applications, the recommended value is $y = 0.01$, which is about as far down as one can go and still retain a positive 2-jet rate. With $\alpha_{\mathrm{s}} = 0.12$, in full second-order QCD (see below), the $2:3:4$ jet composition is then approximately $11 \% : 77 \% : 12 \%$. Note, however, that initial-state QED radiation may occasionally lower the c.m. energy significantly, i.e. increase $\alpha_{\mathrm{s}}$, and thereby bring the 3-jet fraction above unity if $y$ is kept fixed at 0.01 also in those events. Therefore, at PETRA/PEP energies, $y$ values slightly above 0.01 are needed. In addition to the $y$ cut, the program contains a cut on the invariant mass $m_{ij}$ between any two partons, which is typically required to be larger than 2 GeV. This cut corresponds to the actual merging of two nearby parton jets, i.e. where a treatment with two separate partons rather than one would be superfluous in view of the smearing arising from the subsequent fragmentation. Since the cut-off mass scale $\sqrt{y} E_{\mrm{cm}}$ normally is much larger, this additional cut only enters for events at low energies. For massive quarks, the amount of QCD radiation is slightly reduced \cite{Iof78}: \begin{eqnarray} \frac{1}{\sigma_0} \, \frac{\d \sigma}{\d x_1 \, \d x_2} & = & \frac{\alpha_{\mathrm{s}}}{2\pi} \, C_F \, \left\{ \frac{x_1^2 + x_2^2}{(1-x_1)(1-x_2)} - \frac{4 m_{\mathrm{q}}^2}{s} \left( \frac{1}{1-x_1} + \frac{1}{1-x_2} \right) \right. \nonumber \\[1mm] & & - \left. \frac{2 m_{\mathrm{q}}^2}{s} \left( \frac{1}{(1-x_1)^2} + \frac{1}{(1-x_2)^2} \right) - \frac{4 m_{\mathrm{q}}^4}{s^2} \left( \frac{1}{1-x_1} + \frac{1}{1-x_2} \right)^2 \right\} ~. \label{ee:threejMEmass} \end{eqnarray} In addition, the phase space for emission is reduced by the requirement \begin{equation} \frac{(1-x_1)(1-x_2)(1-x_3)}{x_3^2} \geq \frac{m_{\mathrm{q}}^2}{s} ~. \end{equation} For $\b$ quarks at LEP energies, these corrections are fairly small. \subsubsection{4-jet matrix elements} Two new event types are added in second-order QCD, $\e^+\e^- \to \mathrm{q} \overline{\mathrm{q}} \mathrm{g} \mathrm{g}$ and $\e^+\e^- \to \mathrm{q} \overline{\mathrm{q}} \mathrm{q}' \overline{\mathrm{q}}'$. The 4-jet cross section has been calculated by several groups \cite{Ali80a,Gae80,Ell81,Dan82}, which agree on the result. The formulae are too lengthy to be quoted here. In one of the calculations \cite{Ali80a}, quark masses were explicitly included, but {\tsc{Jetset}} only includes the massless expressions, as taken from \cite{Ell81}. Here the angular orientation of the event has been integrated out, so that five independent internal kinematical variables remain. These may be related to the six $y_{ij}$ and the four $y_{ijk}$ variables, $y_{ij} = m_{ij}^2 / s = (p_i + p_j)^2 / s$ and $y_{ijk} = m_{ijk}^2 / s = (p_i + p_j + p_k)^2 / s$, in terms of which the matrix elements are given. The original calculations were for the pure $\gamma$-exchange case; it was recently pointed out \cite{Kni89} that an additional contribution to the $\e^+\e^- \to \mathrm{q} \overline{\mathrm{q}} \mathrm{q}' \overline{\mathrm{q}}'$ cross section arises from the axial part of the $\mathrm{Z}^0$. This term is not included in the program, but fortunately it is finite and small. Whereas the way the string, i.e. the fragmenting colour flux tube, is stretched is uniquely given in $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ event, for $\mathrm{q} \overline{\mathrm{q}} \mathrm{g} \mathrm{g}$ events there are two possibilities: \mbox{$\mathrm{q} - \mathrm{g}_1 - \mathrm{g}_2 - \overline{\mathrm{q}}$} or \mbox{$\mathrm{q} - \mathrm{g}_2 - \mathrm{g}_1 - \overline{\mathrm{q}}$}. A knowledge of quark and gluon colours, obtained by perturbation theory, will uniquely specify the stretching of the string, as long as the two gluons do not have the same colour. The probability for the latter is down in magnitude by a factor $1 / N_C^2 = 1 / 9$. One may either choose to neglect these terms entirely, or to keep them for the choice of kinematical setup, but then drop them at the choice of string drawing \cite{Gus82}. We have adopted the latter procedure. Comparing the two possibilities, differences are typically 10--20\% for a given kinematical configuration, and less for the total 4-jet cross section, so from a practical point of view this is not a major problem. In higher orders, results depend on the renormalization scheme; we will use $\br{\mrm{MS}}$ throughout. In addition to this choice, several possible forms can be chosen for $\alpha_{\mathrm{s}}$, all of which are equivalent to that order but differ in higher orders. We have picked the recommended standard \cite{PDG88} \begin{equation} \label{ee:aS4j} \alpha_{\mathrm{s}}(Q^2) = \frac{12\pi}{(33-2n_f) \, \ln (Q^2 / \Lambda^2_{\br{\mrm{MS}}})} \left\{ 1 - 6 \, \frac{153-19n_f}{(33-2n_f)^2} \, \frac{\ln (\ln ( Q^2 / \Lambda^2_{\br{\mrm{MS}}}))} {\ln ( Q^2 / \Lambda^2_{\br{\mrm{MS}}})} \right\} ~. \end{equation} \subsubsection{Second-order 3-jet matrix elements} As for first order, a full second-order calculation consists both of real parton emission terms and of vertex and propagator corrections. These modify the 3-jet and 2-jet cross sections. Although there was some initial confusion, everybody soon agreed on the size of the loop corrections \cite{Ell81,Ver81,Fab82}. In analytic calculations, the procedure of eq.~(\ref{ee:Obs}), suitably expanded, can therefore be used unambiguously for a well-behaved variable. For Monte Carlo event simulation, it is again necessary to impose some finite jet-resolution criterion. This means that four-parton events which fail the cuts should be reassigned either to the 3-jet or to the 2-jet event class. It is this area that caused quite a lot of confusion in the past \cite{Kun81,Got82,Ali82,Zhu83,Gut84,Gut87,Kra88}, and where full agreement does not exist. Most likely, agreement will never be reached, since there are indeed ambiguous points in the procedure, related to uncertainties on the theoretical side, as follows. For the $y$-cut case, any two partons with an invariant mass $m_{ij}^2 < y E_{\mrm{cm}}^2$ should be recombined into one. If the four-momenta are simply added, the sum will correspond to a parton with a positive mass, namely the original $m_{ij}$. The loop corrections are given in terms of final massless partons, however. In order to perform the (partial) cancellation between the four-parton real and the 3-parton virtual contributions, it is therefore necessary to get rid of the bothersome mass in the four-parton states. Several recombinations are used in practice, which go under names such as `E', `E0', `p' and `p0' \cite{OPA91}. In the `E'-type schemes, the energy of a recombined parton is given by $E_{ij} = E_i + E_j$, and three-momenta may have to be adjusted accordingly. In the `p'-type schemes, on the other hand, three-momenta are added, $\mbf{p}_{ij} = \mbf{p}_i + \mbf{p}_j$, and then energies may have to be adjusted. These procedures result in different 3-jet topologies, and therefore in different second-order differential 3-jet cross sections. Within each scheme, a number of lesser points remain to be dealt with, in particular what to do if a recombination of a nearby parton pair were to give an event with a non-$\mathrm{q}\overline{\mathrm{q}}\mathrm{g}$ flavour structure. {\tsc{Jetset}} contains two alternative second-order 3-jet implementations, GKS and ERT(Zhu). For historical reasons the former is default, but actually the latter is the recommended one today. Other parametrizations have also been made available that run together with {\tsc{Jetset}}, see \cite{Sjo89,Mag89}. The GKS option is based on the GKS \cite{Gut84} calculation, where some of the original mistakes in FKSS \cite{Fab82} have been corrected. The GKS formulae have the advantage of giving the second-order corrections in closed analytic form, as not-too-long functions of $x_1$, $x_2$, and the $y$ cut. However, it is today recognized, also by the authors, that important terms are still missing, and that the matrix elements should therefore not be taken too seriously. The option is thus kept mainly for backwards compatibility. The ERT(Zhu) generator \cite{Zhu83} is based on the ERT matrix elements \cite{Ell81}, with a Monte Carlo recombination procedure suggested by Kunszt \cite{Kun81} and developed by Ali \cite{Ali82}. It has the merit of giving corrections in a convenient, parametrized form. For practical applications, the main limitation is that the corrections are only given for discrete values of the cut-off parameter $y$, namely $y$ = 0.01, 0.02, 0.03, 0.04, and 0.05. The basic approach is the following. Without any loss of generality, the full second-order 3-jet cross section can be written in terms of the `ratio function' $R(X,Y;y)$, defined by \begin{equation} \frac{1}{\sigma_0} \frac{\d \sigma_3^{\mrm{tot}}}{\d X \, \d Y} = \frac{\alpha_{\mathrm{s}}}{\pi} A_0(X,Y) \left\{ 1 + \frac{\alpha_{\mathrm{s}}}{\pi} R(X,Y;y) \right\} ~, \label{ee:Zhupar} \end{equation} where $X = x_1 - x_2 = x_{\mathrm{q}} - x_{\overline{\mathrm{q}}}$, $Y = x_3 = x_g$, $\sigma_0$ is the lowest-order hadronic cross section, and $A_0(X,Y)$ the standard first-order 3-jet cross section, cf. eq.~(\ref{ee:ME3j}). By Monte Carlo integration, the value of $R(X,Y;y)$ is evaluated in bins of $(X,Y)$, and the result parametrized by a simple function $F(X,Y;y)$. In order to obtain the second-order 3-jet rate, a small cut $y_0 = 10^{-7}$ was introduced. It was assumed that four-parton events which fail this cut can be (partly) cancelled analytically against the virtual 3-jet events, to give a net `regularized virtual' contribution to the 3-jet rate. For a given choice of $y$ cut, in the physical range $y \gg y_0$, an additional `soft' contribution comes from four-parton events which survive the $y_0$ cut but fail the $y$ one. A large sample (9\,000\,000) of four-parton events was generated inside the $y_0$ cut region. For events which failed the more stringent $y$ cuts, the parton pair with the smallest invariant mass was recombined into an effective jet, using the `p0' recombination scheme. This means that the individual three-momenta were added, $\mbf{p}_{ij} = \mbf{p}_i + \mbf{p}_j$, the mass of the recombined pair was set to zero for the calculation of energy, $E_{ij} = |\mbf{p}_i + \mbf{p}_j|$, and finally all four-momenta were rescaled by a common factor so as to preserve the correct c.m. frame energy. In calculating the ${\cal O}(\alpha_{\mathrm{s}}^2)$ correction functions, care was taken to maintain the flavour signature of the jets in the recombination process. A quark and a gluon were recombined into a quark with the same flavour as the original quark, two gluons were recombined to form a gluon, etc. In some cases the three jets of the final state were not in the standard $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ configuration. The probability for this to happen corresponded to less than 0.5\% of the total cross section, even for the most stringent cuts used. For these non-$\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ final states, the assignment of $\mathrm{q}$, $\overline{\mathrm{q}}$ and $\mathrm{g}$ was done at random. The sum of `regularized virtual' (1\,000\,000 3-jet events were generated, with evaluated second-order weights) and `soft' corrections, normalized to the first-order 3-jet cross section, was tabulated in the $(X,Y)$ plane, using bins of size $0.05 \times 0.05$. This estimated $R$-function behaviour was then fit with a 12-parameter function $F$, \begin{eqnarray} F(X,Y;y)& = & p_1 + p_2 X^2 + p_3 X^4 + (p_4+p_5 X^2)Y+ (p_6+p_7X^2)Y^2 + \nonumber\\ & & (p_8+p_9X^2)Y^3 + p_{10}/(X^2-Y^2) + p_{11}/(1-Y) + p_{12}/Y ~. \label{ee:ZhuFun} \end{eqnarray} The parameters $p_i$ are reproduced in \cite{Sjo89}. \subsubsection{The matrix-element event generator scheme} The program contains parametrizations, separately, of the total first-order 3-jet rate, the total second-order 3-jet rate, and the total 4-jet rate, all as functions of $y$ (with $\alpha_{\mathrm{s}}$ as a separate prefactor). These parametrizations have been obtained as follows: \begin{Itemize} \item The first-order 3-jet matrix element is almost analytically integrable; some small finite pieces were obtained by a truncated series expansion of the relevant integrand. \item The GKS second-order 3-jet matrix elements were integrated for 40 different $y$-cut values, evenly distributed in $\ln y$ between a smallest value $y = 0.001$ and the kinematical limit $y = 1/3$. For each $y$ value, 250\,000 phase-space points were generated, evenly in $\d \ln (1-x_i) = \d x_i/(1-x_i)$, $i = 1,2$, and the second-order 3-jet rate in the point evaluated. The properly normalized sum of weights in each of the 40 $y$ points were then fitted to a polynomial in $\ln(y^{-1}-2)$. For the ERT(Zhu) matrix elements the parametrizations in eq.~(\ref{ee:ZhuFun}) were used to perform a corresponding Monte Carlo integration for the five $y$ values available. \item The 4-jet rate was integrated numerically, separately for $\mathrm{q} \overline{\mathrm{q}} \mathrm{g} \mathrm{g}$ and $\mathrm{q} \overline{\mathrm{q}} \mathrm{q}' \overline{\mathrm{q}}'$ events, by generating large samples of 4-jet phase-space points within the boundary $y = 0.001$. Each point was classified according to the actual minimum $y$ between any two partons. The same events could then be used to update the summed weights for 40 different counters, corresponding to $y$ values evenly distributed in $\ln y$ between $y = 0.001$ and the kinematical limit $y = 1/6$. In fact, since the weight sums for large $y$ values only received contributions from few phase-space points, extra (smaller) subsamples of events were generated with larger $y$ cuts. The summed weights, properly normalized, were then parametrized in terms of polynomials in $\ln(y^{-1} - 5)$. Since it turned out to be difficult to obtain one single good fit over the whole range of $y$ values, different parametrizations are used above and below $y=0.018$. As originally given, the $\mathrm{q} \overline{\mathrm{q}} \mathrm{q}' \overline{\mathrm{q}}'$ parametrization only took into account four $\mathrm{q}'$ flavours, i.e. secondary $\b \overline{\mathrm{b}}$ pairs were not generated, but this has been corrected for LEP. \end{Itemize} In the generation stage, each event is treated on its own, which means that the $\alpha_{\mathrm{s}}$ and $y$ values may be allowed to vary from event to event. The main steps are the following. \begin{Enumerate} \item The $y$ value to be used in the current event is determined. If possible, this is the value given by you, but additional constraints exist from the validity of the parametrizations ($y \geq 0.001$ for GKS, $0.01 \leq y \leq 0.05$ for ERT(Zhu)) and an extra (user-modifiable) requirement of a minimum absolute invariant mass between jets (which translates into varying $y$ cuts due to the effects of initial-state QED radiation). \item The $\alpha_{\mathrm{s}}$ value is calculated. \item For the $y$ and $\alpha_{\mathrm{s}}$ values given, the relative two/three/4-jet composition is determined. This is achieved by using the parametrized functions of $y$ for 3- and 4-jet rates, multiplied by the relevant number of factors of $\alpha_{\mathrm{s}}$. In ERT(Zhu), where the second-order 3-jet rate is available only at a few $y$ values, intermediate results are obtained by linear interpolation in the ratio of second-order to first-order 3-jet rates. The 3-jet and 4-jet rates are normalized to the analytically known second-order total event rate, i.e. divided by $R_{\mrm{QCD}}$ of eq.~(\ref{ee:RQCD}). Finally, the 2-jet rate is obtained by conservation of total probability. \item If the combination of $y$ and $\alpha_{\mathrm{s}}$ values is such that the total 3- plus 4-jet fraction is larger than unity, i.e. the remainder 2-jet fraction negative, the $y$-cut value is raised (for that event), and the process is started over at point 3. \item The choice is made between generating a 2-, 3- or 4-jet event, according to the relative probabilities. \item For the generation of 4-jets, it is first necessary to make a choice between $\mathrm{q} \overline{\mathrm{q}} \mathrm{g} \mathrm{g}$ and $\mathrm{q} \overline{\mathrm{q}} \mathrm{q}' \overline{\mathrm{q}}'$ events, according to the relative (parametrized) total cross sections. A phase-space point is then selected, and the differential cross section at this point is evaluated and compared with a parametrized maximum weight. If the phase-space point is rejected, a new one is selected, until an acceptable 4-jet event is found. \item For 3-jets, a phase-space point is first chosen according to the first-order cross section. For this point, the weight \begin{equation} W(x_1,x_2;y) = 1 + \frac{\alpha_{\mathrm{s}}}{\pi} R(x_1,x_2;y) \label{ee:WTJS} \end{equation} is evaluated. Here $R(x_1,x_2;y)$ is analytically given for GKS \cite{Gut84}, while it is approximated by the parametrization $F(X,Y;y)$ of eq.~(\ref{ee:ZhuFun}) for ERT(Zhu). Again, linear interpolation of $F(X,Y;y)$ has to be applied for intermediate $y$ values. The weight $W$ is compared with a maximum weight \begin{equation} W_{\mathrm{max}}(y) = 1 + \frac{\alpha_{\mathrm{s}}}{\pi} R_{\mathrm{max}}(y) ~, \end{equation} which has been numerically determined beforehand and suitably parametrized. If the phase-space point is rejected, a new point is generated, etc. \item Massive matrix elements are not available in {\tsc{Jetset}} for second-order QCD (but are in the first-order option). However, if a 3- or 4-jet event determined above falls outside the phase-space region allowed for massive quarks, the event is rejected and reassigned to be a 2-jet event. (The way the $y_{ij}$ and $y_{ijk}$ variables of 4-jet events should be interpreted for massive quarks is not even unique, so some latitute has been taken here to provide a reasonable continuity from 3-jet events.) This procedure is known not to give the expected full mass suppression, but is a reasonable first approximation. \item Finally, if the event is classified as a 2-jet event, either because it was initially so assigned, or because it failed the massive phase-space cuts for 3- and 4-jets, the generation of 2-jets is trivial. \end{Enumerate} \subsubsection{Optimized perturbation theory} Theoretically, it turns out that the second-order corrections to the 3-jet rate are large. It is therefore not unreasonable to expect large third-order corrections to the 4-jet rate. Indeed, the experimental 4-jet rate is much larger than second order predicts (when fragmentation effects have been folded in), if $\alpha_{\mathrm{s}}$ is determined based on the 3-jet rate \cite{Sjo84a,JAD88}. The only consistent way to resolve this issue is to go ahead and calculate the full next order. This is a tough task, however, so people have looked at possible shortcuts. For example, one can try to minimize the higher-order contributions by a suitable choice of the renormalization scale \cite{Ste81} --- `optimized perturbation theory'. This is equivalent to a different choice for the $Q^2$ scale in $\alpha_{\mathrm{s}}$, a scale which is not unambiguous anyway. Indeed the standard value $Q^2 = s = E_{\mrm{cm}}^2$ is larger than the natural physical scale of gluon emission in events, given that most gluons are fairly soft. One could therefore pick another scale, $Q^2 = f s$, with $f < 1$. The ${\cal O}(\alpha_{\mathrm{s}})$ 3-jet rate would be increased by such a scale change, and so would the number of 4-jet events, including those which collapse into 3-jet ones. The loop corrections depend on the $Q^2$ scale, however, and compensate the changes above by giving a larger negative contribution to the 3-jet rate. The possibility of picking an optimized scale $f$ is implemented as follows \cite{Sjo89}. Assume that the differential 3-jet rate at scale $Q^2 = s$ is given by the expression \begin{equation} R_3 = r_1 \alpha_{\mathrm{s}} + r_2 \alpha_{\mathrm{s}}^2 ~, \end{equation} where $R_3$, $r_1$ and $r_2$ are functions of the kinematical variables $x_1$ and $x_2$ and the $y$ cut, as described above. When the coupling is chosen at a different scale, $Q'^2 = f s$, the 3-jet rate has to be changed to \begin{equation} R_3' = r_1' \alpha_{\mathrm{s}}' + r_2 \alpha_{\mathrm{s}}'^2 ~, \end{equation} where $r_1' = r_1$, \begin{equation} r_2' = r_2 + r_1 \frac{33-2n_f}{12\pi} \ln f ~, \label{ee:r2optim} \end{equation} and $\alpha_{\mathrm{s}}' = \alpha_{\mathrm{s}}(fs)$. Since we only have the Born term for 4-jets, here the effects of a scale change come only from the change in the coupling constant. Finally, the 2-jet cross section can still be calculated from the difference between the total cross section and the 3- and 4-jet cross sections. If an optimized scale is used in the program, the default value is $f=0.002$, which is favoured by the studies in ref. \cite{Bet89}. (In fact, it is also possible to use a correspondingly optimized $R_{\mrm{QCD}}$ factor, eq.~(\ref{ee:RQCD}), but then the corresponding $f$ is chosen independently and much closer to unity.) The success of describing the jet rates should not hide the fact that one is dabbling in (educated, hopefully) guesswork, and that any conclusions based on this method have to be taken with a pinch of salt. One special problem associated with the use of optimized perturbation theory is that the differential 3-jet rate may become negative over large regions of the $(x_1, x_2)$ phase space. This problem already exists, at least in principle, even for a scale $f = 1$, since $r_2$ is not guaranteed to be positive definite. Indeed, depending on the choice of $y$ cut, $\alpha_{\mathrm{s}}$ value and recombination scheme, one may observe a small region of negative differential 3-jet rate for the full second-order expression. This region is centred around $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ configurations, where the $\mathrm{q}$ and $\overline{\mathrm{q}}$ are close together in one hemisphere and the $\mathrm{g}$ is alone in the other, i.e. $x_1 \approx x_2 \approx 1/2$. It is well understood why second-order corrections should be negative in this region \cite{Dok89}: the $\mathrm{q}$ and $\overline{\mathrm{q}}$ of a $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ state are in a relative colour octet state, and thus the colour force between them is repulsive, which translates into a negative second-order term. However, as $f$ is decreased below unity, $r_2'$ receives a negative contribution from the $\ln f$ term, and the region of negative differential cross section has a tendency to become larger, also after taking into account related changes in $\alpha_{\mathrm{s}}$. In an event-generator framework, where all events are supposed to come with unit weight, it is clearly not possible to simulate negative cross sections. What happens in the program is therefore that no 3-jet events at all are generated in the regions of negative differential cross section, and that the 3-jet rate in regions of positive cross sections is reduced by a constant factor, chosen so that the total number of 3-jet events comes out as it should. This is a consequence of the way the program works, where it is first decided what kind of event to generate, based on integrated 3-jet rates in which positive and negative contributions are added up with sign, and only thereafter the kinematics is chosen. Based on our physics understanding of the origin of this negative cross section, the approach adopted is as sensible as any, at least to that order in perturbation theory (what one might strive for is a properly exponentiated description of the relevant region). It can give rise to funny results for low $f$ values, however, as observed by OPAL \cite{OPA92} for the energy--energy correlation asymmetry. \subsubsection{Angular orientation} While pure $\gamma$ exchange gives a simple $1 + \cos^2\theta$ distribution for the $\mathrm{q}$ (and $\overline{\mathrm{q}}$) direction in $\mathrm{q} \overline{\mathrm{q}}$ events, $\mathrm{Z}^0$ exchange and $\gamma^* / \Z^0$ interference results in a forward--backward asymmetry. If one introduces \begin{eqnarray} h'_{\mathrm{f}}(s) & = & 2 e_{\mathrm{e}} \left\{ a_{\mathrm{e}} (1 - P_{\mrm{L}}^+ P_{\mrm{L}}^-) - v_{\mathrm{e}} (P_{\mrm{L}}^- - P_{\mrm{L}}^+) \right\} \, \Re\chi(s) e_{\mathrm{f}} a_{\mathrm{f}} \nonumber \\ & & + \, \left\{ 2 v_{\mathrm{e}} a_{\mathrm{e}} (1 - P_{\mrm{L}}^+ P_{\mrm{L}}^-) - (v_{\mathrm{e}}^2 + a_{\mathrm{e}}^2) (P_{\mrm{L}}^- - P_{\mrm{L}}^+) \right\} \, |\chi(s)|^2 \, v_{\mathrm{f}} a_{\mathrm{f}} ~, \end{eqnarray} then the angular distribution of the quark is given by \begin{equation} \frac{\d \sigma}{\d (\cos\theta_{\mathrm{f}})} \propto h_{\mathrm{f}}(s)(1 + \cos^2\theta_{\mathrm{f}}) + 2 h'_{\mathrm{f}}(s) \cos\theta_{\mathrm{f}} ~. \end{equation} The angular orientation of a 3- or 4-jet event may be described in terms of three angles $\chi$, $\theta$ and $\varphi$; for 2-jet events only $\theta$ and $\varphi$ are necessary. From a standard orientation, with the $\mathrm{q}$ along the $+z$ axis and the $\overline{\mathrm{q}}$ in the $xz$ plane with $p_x > 0$, an arbitrary orientation may be reached by the rotations $+\chi$ in azimuthal angle, $+\theta$ in polar angle, and $+\varphi$ in azimuthal angle, in that order. Differential cross sections, including QFD effects and arbitrary beam polarizations have been given for 2- and 3-jet events in refs. \cite{Ols80,Sch80}. We use the formalism of ref. \cite{Ols80}, with $\chi \to \pi - \chi$ and $\varphi^- \to - (\varphi + \pi/2)$. The resulting formulae are tedious, but straightforward to apply, once the internal jet configuration has been chosen. 4-jet events are approximated by 3-jet ones, by joining the two gluons of a $\mathrm{q} \overline{\mathrm{q}} \mathrm{g} \mathrm{g}$ event and the $\mathrm{q}'$ and $\overline{\mathrm{q}}'$ of a $\mathrm{q} \overline{\mathrm{q}} \mathrm{q}' \overline{\mathrm{q}}'$ event into one effective jet. This means that some angular asymmetries are neglected \cite{Ali80a}, but that weak effects are automatically included. It is assumed that the second-order 3-jet events have the same angular orientation as the first-order ones, some studies on this issue may be found in \cite{Kor85}. Further, the formulae normally refer to the massless case; only for the QED 2- and 3-jet cases are mass corrections available. The main effect of the angular distribution of multijet events is to smear the lowest-order result, i.e. to reduce any anisotropies present in 2-jet systems. In the parton-shower option of the program, only the initial $\mathrm{q} \overline{\mathrm{q}}$ axis is determined. The subsequent shower evolution then {\it de facto} leads to a smearing of the jet axis, although not necessarily in full agreement with the expectations from multijet matrix-element treatments. \subsubsection{Initial-state radiation} Initial-state photon radiation has been included using the formalism of ref. \cite{Ber82}. Here each event contains either no photon or one, i.e. it is a first-order non-exponentiated description. The main formula for the hard radiative photon cross section is \begin{equation} \frac{\d \sigma}{\d x_{\gamma}} = \frac{\alpha_{\mathrm{em}}}{\pi} \, \left( \ln\frac{s}{m_{\mathrm{e}}^2} -1 \right) \, \frac{1 + (1-x_{\gamma})^2}{x_{\gamma}} \, \sigma_0 (\hat{s}) ~, \end{equation} where $x_{\gamma}$ is the photon energy fraction of the beam energy, $\hat{s} = (1-x_{\gamma}) s$ is the squared reduced hadronic c.m. energy, and $\sigma_0$ is the ordinary annihilation cross section at the reduced energy. In particular, the selection of jet flavours should be done according to expectations at the reduced energy. The cross section is divergent both for $x_{\gamma} \to 1$ and $x_{\gamma} \to 0$. The former is related to the fact that $\sigma_0$ has a $1/\hat{s}$ singularity (the real photon pole) for $\hat{s} \to 0$. An upper cut on $x_{\gamma}$ can here be chosen to fit the experimental setup. The latter is a soft photon singularity, which is to be compensated in the no-radiation cross section. A requirement $x_{\gamma} > 0.01$ has therefore been chosen so that the hard-photon fraction is smaller than unity. In the total cross section, effects from photons with $x_{\gamma} < 0.01$ are taken into account, together with vertex and vacuum polarization corrections (hadronic vacuum polarizations using a simple parametrization of the more complicated formulae of ref. \cite{Ber82}). The hard photon spectrum can be integrated analytically, for the full $\gamma^* / \Z^0$ structure including interference terms, provided that no new flavour thresholds are crossed and that the $R_{\mrm{QCD}}$ term in the cross section can be approximated by a constant over the range of allowed $\hat{s}$ values. In fact, threshold effects can be taken into account by standard rejection techniques, at the price of not obtaining the exact cross section analytically, but only by an effective Monte Carlo integration taking place in parallel with the ordinary event generation. In addition to $x_{\gamma}$, the polar angle $\theta_{\gamma}$ and azimuthal angle $\varphi_{\gamma}$ of the photons are also to be chosen. Further, for the orientation of the hadronic system, a choice has to be made whether the photon is to be considered as having been radiated from the $\mathrm{e}^+$ or from the $\mathrm{e}^-$. Final-state photon radiation, as well as interference between initial- and final-state radiation, has been left out of this treatment. The formulae for $\e^+\e^- \to \mu^+ \mu^-$ cannot be simply taken over for the case of outgoing quarks, since the quarks as such only live for a short while before turning into hadrons. Another simplification in our treatment is that effects of incoming polarized $\mathrm{e}^{\pm}$ beams have been completely neglected, i.e. neither the effective shift in azimuthal distribution of photons nor the reduction in polarization is included. The polarization parameters of the program are to be thought of as the effective polarization surviving after initial-state radiation. \subsubsection{Alternative matrix elements} The program contains two sets of `toy model' matrix elements, one for an Abelian vector gluon model and one for a scalar gluon model. Clearly both of these alternatives are already excluded by data, and are anyway not viable alternatives for a consistent theory of strong interactions. They are therefore included more as references to show how well the characteristic features of QCD can be measured experimentally. Second-order matrix elements are available for the Abelian vector gluon model. These are easily obtained from the standard QCD matrix elements by a substitution of the Casimir group factors: $C_F = 4/3 \to 1$, $N_C = 3 \to 0$, and $T_R = n_{\mathrm{f}}/2 \to 3 n_{\mathrm{f}}$. First-order matrix elements contain only $C_F$; therefore the standard first-order QCD results may be recovered by a rescaling of $\alpha_{\mathrm{s}}$ by a factor $4/3$. In second order the change of $N_C$ to 0 means that $\mathrm{g} \to \mathrm{g}\g$ couplings are absent from the Abelian model, while the change of $T_R$ corresponds to an enhancement of the $\mathrm{g} \to \mathrm{q}'\overline{\mathrm{q}}'$ coupling, i.e. to an enhancement of the $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ 4-jet event rate. The second-order corrections to the 3-jet rate turn out to be strongly negative --- if $\alpha_{\mathrm{s}}$ is fitted to get about the right rate of 4-jet events, the predicted differential 3-jet rate is negative almost everywhere in the $(x_1, x_2)$ plane. Whether this unphysical behaviour would be saved by higher orders is unclear. It has been pointed out that the rate can be made positive by a suitable choice of scale, since $\alpha_{\mathrm{s}}$ runs in opposite directions in an Abelian model and in QCD \cite{Bet89}. This may be seen directly from eq. (\ref{ee:r2optim}), where the term $33 = 11 N_C$ is absent in the Abelian model, and therefore the scale-dependent term changes sign. In the program, optimized scales have not been implemented for this toy model. Therefore the alternatives provided for you are either to generate only 4-jet events, or to neglect second-order corrections to the 3-jet rate, or to have the total 3-jet rate set vanishing (so that only 2- and 4-jet events are generated). Normally we would expect the former to be the one of most interest, since it is in angular (and flavour) distributions of 4-jet events that the structure of QCD can be tested. Also note that the `correct' running of $\alpha_{\mathrm{s}}$ is not included; you are expected to use the option where $\alpha_{\mathrm{s}}$ is just given as a constant number. The scalar gluon model is even more excluded than the Abelian vector one, since differences appear already in the 3-jet matrix element \cite{Lae80}: \begin{equation} \frac{\d \sigma}{\d x_1 \, \d x_2} \propto \frac{x_3^2}{(1-x_1)(1-x_2)} \end{equation} when only $\gamma$ exchange is included. The axial part of the $\mathrm{Z}^0$ gives a slightly different shape; this is included in the program but does not make much difference. The angular orientation does include the full $\gamma^* / \Z^0$ interference \cite{Lae80}, but the main interest is in the 3-jet topology as such \cite{Ell79}. No higher-order corrections are included. It is recommended to use the option of a fixed $\alpha_{\mathrm{s}}$ also here, since the correct running is not available. \subsection{Decays of Onia Resonances} \label{ss:oniadecays} Many different possibilities are open for the decay of heavy $J^{PC} = 1^{--}$ onia resonances. Of special interest are the decays into three gluons or two gluons plus a photon, since these offer unique possibilities to study a `pure sample' of gluon jets. A routine for this purpose is included in the program. It was written at a time where the expectations were to find toponium at PETRA energies. If, as now seems likely, the top mass is above 100~GeV, weak decays will dominate, to the extent that the top quark will decay weakly even before a bound toponium state is formed, and thus the routine will be of no use for top. The charm system, on the other hand, is far too low in mass for a jet language to be of any use. The only application is therefore likely to be for $\Upsilon$, which unfortunately also is on the low side in mass. The matrix element for $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{g} \mathrm{g} \mathrm{g}$ is (in lowest order) \cite{Kol78} \begin{equation} \frac{1}{\sigma_{\mathrm{g} \mathrm{g} \mathrm{g}}} \frac{\d \sigma_{\mathrm{g} \mathrm{g} \mathrm{g}}}{\d x_1 \, \d x_2} = \frac{1}{\pi^2 - 9} \left\{ \left( \frac{1-x_1}{x_2 x_3} \right)^2 + \left( \frac{1-x_2}{x_1 x_3} \right)^2 + \left( \frac{1-x_3}{x_1 x_2} \right)^2 \right\} ~, \label{ee:Upsilondec} \end{equation} where, as before, $x_i = 2 E_i / E_{\mrm{cm}}$ in the c.m. frame of the event. This is a well-defined expression, without the kind of singularities encountered in the $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ matrix elements. In principle, no cuts at all would be necessary, but for reasons of numerical simplicity we implement a $y$ cut as for continuum jet production, with all events not fulfilling this cut considered as (effective) $\mathrm{g} \mathrm{g}$ events. For $\mathrm{g} \mathrm{g} \mathrm{g}$ events, each $\mathrm{g} \mathrm{g}$ invariant mass is required to be at least 2 GeV. Another process is $\mathrm{q} \overline{\mathrm{q}} \to \gamma \mathrm{g} \mathrm{g}$, obtained by replacing a gluon in $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{g} \mathrm{g} \mathrm{g}$ by a photon. This process has the same normalized cross section as the one above, if e.g. $x_1$ is taken to refer to the photon. The relative rate is \cite{Kol78} \begin{equation} \frac{\sigma_{\gamma \mathrm{g} \mathrm{g}}}{\sigma_{\mathrm{g} \mathrm{g} \mathrm{g}}} = \frac{36}{5} \, \frac{e_{\mathrm{q}}^2 \, \alpha_{\mathrm{em}}} {\alpha_{\mathrm{s}}(Q^2)} ~. \end{equation} Here $e_{\mathrm{q}}$ is the charge of the heavy quark, and the scale in $\alpha_{\mathrm{s}}$ has been chosen as the mass of the onium state. If the mass of the recoiling $\mathrm{g} \mathrm{g}$ system is lower than some cut-off (by default 2 GeV), the event is rejected. In the present implementation the angular orientation of the $\mathrm{g} \mathrm{g} \mathrm{g}$ and $\gamma \mathrm{g} \mathrm{g}$ events is given for the $\e^+\e^- \to \gamma^* \to$ onium case \cite{Kol78} (optionally with beam polarization effects included), i.e. weak effects have not been included, since they are negligible at around 10~GeV. It is possible to start a perturbative shower evolution from either of the two states above. However, for $\Upsilon$ the phase space for additional evolution is so constrained that not much is to be gained from that. We therefore do not recommend this possibility. The shower generation machinery, when starting up from a $\gamma \mathrm{g} \mathrm{g}$ configuration, is constructed such that the photon energy is not changed. This means that there is currently no possibility to use showers to bring the theoretical photon spectrum in better agreement with the experimental one. In string fragmentation language, a $\mathrm{g} \mathrm{g} \mathrm{g}$ state corresponds to a closed string triangle with the three gluons at the corners. As the partons move apart from a common origin, the string triangle expands. Since the photon does not take part in the fragmentation, the $\gamma \mathrm{g} \mathrm{g}$ state corresponds to a double string running between the two gluons. \subsection{Routines and Common Block Variables} \label{ss:eeroutines} \subsubsection{$\e^+\e^-$ continuum event generation} The only routine a normal user will call to generate $\e^+\e^-$ continuum events is \ttt{LUEEVT}. The other routines listed below, as well as \ttt{LUSHOW} (see section \ref{ss:showrout}), are called by \ttt{LUEEVT}. \drawbox{CALL LUEEVT(KFL,ECM)}\label{p:LUEEVT} \begin{entry} \itemc{Purpose:} to generate a complete event $\e^+\e^- \to \gamma^* / \Z^0 \to \mathrm{q}\overline{\mathrm{q}} \to$ parton shower $\to$ hadrons according to QFD and QCD cross sections. As an alternative to parton showers, second-order matrix elements are available for $\mathrm{q}\overline{\mathrm{q}} + \mathrm{q}\overline{\mathrm{q}}\mathrm{g} + \mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g + \mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ production. \iteme{KFL :} flavour of events generated. \begin{subentry} \iteme{= 0 :} mixture of all allowed flavours according to relevant probabilities. \iteme{= 1 - 8 :} primary quarks are only of the specified flavour \ttt{KFL}. \end{subentry} \iteme{ECM :} total c.m. energy of system. \itemc{Remark:} Each call generates one event, which is independent of preceding ones, with one exception, as follows. If radiative corrections are included, the shape of the hard photon spectrum is recalculated only with each \ttt{LUXTOT} call, which normally is done only if \ttt{KFL}, \ttt{ECM} or \ttt{MSTJ(102)} is changed. A change of e.g. the $\mathrm{Z}^0$ mass in mid-run has to be followed either by a user call to \ttt{LUXTOT} or by an internal call forced e.g. by putting \ttt{MSTJ(116)=3}. \end{entry} \boxsep \begin{entry} \iteme{SUBROUTINE LUXTOT(KFL,ECM,XTOT) :}\label{p:LUXTOT} to calculate the total hadronic cross section, including quark thresholds, weak, beam polarization, and QCD effects and radiative corrections. In the process, variables necessary for the treatment of hard photon radiation are calculated and stored. \begin{subentry} \iteme{KFL, ECM :} as for \ttt{LUEEVT}. \iteme{XTOT :} the calculated total cross section in nb. \end{subentry} \iteme{SUBROUTINE LURADK(ECM,MK,PAK,THEK,PHIK,ALPK) :}\label{p:LURADK} to describe initial-state hard $\gamma$ radiation. \iteme{SUBROUTINE LUXKFL(KFL,ECM,ECMC,KFLC) :}\label{p:LUXKFL} to generate the primary quark flavour in case this is not specified by the user. \iteme{SUBROUTINE LUXJET(ECM,NJET,CUT) :}\label{p:LUXJET} to determine the number of jets (2, 3 or 4) to be generated within the kinematically allowed region (characterized by \ttt{CUT} $= y_{\mrm{cut}}$) in the matrix-element approach; to be chosen such that all probabilities are between 0 and 1. \iteme{SUBROUTINE LUX3JT(NJET,CUT,KFL,ECM,X1,X2) :}\label{p:LUX3JT} to generate the internal momentum variables of a 3-jet event, $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}$, according to first- or second-order QCD matrix elements. \iteme{SUBROUTINE LUX4JT(NJET,CUT,KFL,ECM,KFLN,X1,X2,X4,X12,X14) :}% \label{p:LUX4JT} to generate the internal momentum variables for a 4-jet event, $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g$ or $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$, according to second-order QCD matrix elements. \iteme{SUBROUTINE LUXDIF(NC,NJET,KFL,ECM,CHI,THE,PHI) :}\label{p:LUXDIF} to describe the angular orientation of the jets. In first-order QCD the complete QED or QFD formulae are used; in second order 3-jets are assumed to have the same orientation as in first, and 4-jets are approximated by 3-jets. \end{entry} \subsubsection{A routine for onium decay} In \ttt{LUONIA} we have implemented the decays of heavy onia resonances into three gluons or two gluons plus a photon, which are the dominant non-background-like decays of $\Upsilon$. \drawbox{CALL LUONIA(KFL,ECM)}\label{p:LUONIA} \begin{entry} \itemc{Purpose:} to simulate the process $\e^+\e^- \to \gamma^* \to 1^{--}$ onium resonance $\to (\mathrm{g}\g\mathrm{g}$ or $\mathrm{g}\g\gamma) \to$ shower $\to$ hadrons. \iteme{KFL :} the flavour of the quark giving rise to the resonance. \begin{subentry} \iteme{= 0 :} generate $\mathrm{g}\g\mathrm{g}$ events alone. \iteme{= 1 - 8 :} generate $\mathrm{g}\g\mathrm{g}$ and $\mathrm{g}\g\gamma$ events in mixture determined by the squared charge of flavour \ttt{KFL}. Normally \ttt{KFL=} 5 or 6. \end{subentry} \iteme{ECM :} total c.m. energy of system. \end{entry} \subsubsection{Common block variables} The status codes and parameters relevant for the $\e^+\e^-$ routines are found in the common block \ttt{LUDAT1}. This common block also contains more general status codes and parameters, described elsewhere. \drawbox{COMMON/LUDAT1/MSTU(200),PARU(200),MSTJ(200),PARJ(200)} \begin{entry} \itemc{Purpose:} to give access to a number of status codes and parameters regulating the performance of the $\e^+\e^-$ event generation routines. \iteme{MSTJ(101) :}\label{p:MSTJ101} (D=5) gives the type of QCD corrections used for continuum events. \begin{subentry} \iteme{= 0 :} only $\mathrm{q}\overline{\mathrm{q}}$ events are generated. \iteme{= 1 :} $\mathrm{q}\overline{\mathrm{q}} + \mathrm{q}\overline{\mathrm{q}}\mathrm{g}$ events are generated according to first-order QCD. \iteme{= 2 :} $\mathrm{q}\overline{\mathrm{q}} + \mathrm{q}\overline{\mathrm{q}}\mathrm{g} + \mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g + \mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ events are generated according to second-order QCD. \iteme{= 3 :} $\mathrm{q}\overline{\mathrm{q}} + \mathrm{q}\overline{\mathrm{q}}\mathrm{g} + \mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g + \mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ events are generated, but without second-order corrections to the 3-jet rate. \iteme{= 5 :} a parton shower is allowed to develop from an original $\mathrm{q}\overline{\mathrm{q}}$ pair, see \ttt{MSTJ(40) - MSTJ(50)} for details. \iteme{= -1 :} only $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}$ events are generated (within same matrix-element cuts as for \ttt{=1}). Since the change in flavour composition from mass cuts or radiative corrections is not taken into account, this option is not intended for quantitative studies. \iteme{= -2 :} only $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g$ and $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ events are generated (as for \ttt{=2}). The same warning as for \ttt{=-1} applies. \iteme{= -3 :} only $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g$ events are generated (as for \ttt{=2}). The same warning as for \ttt{=-1} applies. \iteme{= -4 :} only $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ events are generated (as for \ttt{=2}). The same warning as for \ttt{=-1} applies. \itemc{Note 1:} \ttt{MSTJ(101)} is also used in \ttt{LUONIA}, with \iteme{$\leq$ 4 :} $\mathrm{g}\g\mathrm{g} + \gamma\mathrm{g}\g$ events are generated according to lowest-order matrix elements. \iteme{$\geq$ 5 :} a parton shower is allowed to develop from the original $\mathrm{g}\g\mathrm{g}$ or $\mathrm{g}\g\gamma$ configuration, see \ttt{MSTJ(40) - MSTJ(50)} for details. \itemc{Note 2:} The default values of fragmentation parameters have been chosen to work well with the default parton-shower approach above. If any of the other options are used, or if the parton shower is used in non-default mode, it may be necessary to retune fragmentation parameters. As an example, we note that the second-order matrix-element approach (\ttt{MSTJ(101)=2}) at PETRA/PEP energies gives a better description when the $a$ and $b$ parameters of the symmetric fragmentation function are set to $a =$\ttt{PARJ(41)=1}, $b =$\ttt{PARJ(42)=0.7}, and the width of the transverse momentum distribution to $\sigma =$\ttt{PARJ(21)=0.40}. In principle, one also ought to change the joining parameter to \ttt{PARJ(33)=PARJ(35)=1.1} to preserve a flat rapidity plateau, but if this should be forgotten, it does not make too much difference. For applications at TRISTAN or LEP, one must expect to have to change the matrix-element approach parameters even more, to make up for additional soft gluon effects not covered in this approach. \end{subentry} \iteme{MSTJ(102) :} (D=2) inclusion of weak effects ($\mathrm{Z}^0$ exchange) for flavour production, angular orientation, cross sections and initial-state photon radiation in continuum events. \begin{subentry} \iteme{= 1 :} QED, i.e. no weak effects are included. \iteme{= 2 :} QFD, i.e. including weak effects. \iteme{= 3 :} as \ttt{=2}, but at initialization in \ttt{LUXTOT} the $\mathrm{Z}^0$ width is calculated from $\sin^2 \! \theta_W$, $\alpha_{\mathrm{em}}$ and $\mathrm{Z}^0$ and quark masses (including bottom and top threshold factors for \ttt{MSTJ(103)} odd), assuming three full generations, and the result is stored in \ttt{PARJ(124)}. \end{subentry} \iteme{MSTJ(103) :} (D=7) mass effects in continuum matrix elements, in the form \ttt{MSTJ(103)} $= M_1 + 2M_2 + 4M_3$, where $M_i = 0$ if no mass effects and $M_i = 1$ if mass effects should be included. Here; \begin{subentry} \iteme{$M_1$ :} threshold factor for new flavour production according to QFD result; \iteme{$M_2$ :} gluon emission probability (only applies for \ttt{|MSTJ(101)|}$\leq 1$, otherwise no mass effects anyhow); \iteme{$M_3$ :} angular orientation of event (only applies for \ttt{|MSTJ(101)|}$\leq 1$ and \ttt{MSTJ(102)=1}, otherwise no mass effects anyhow). \end{subentry} \iteme{MSTJ(104) :} (D=5) number of allowed flavours, i.e. flavours that can be produced in a continuum event if the energy is enough. A change to 6 makes top production allowed above the threshold, etc. Note that in $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ events only the first five flavours are allowed in the secondary pair, produced by a gluon breakup. \iteme{MSTJ(105) :} (D=1) fragmentation and decay in \ttt{LUEEVT} and \ttt{LUONIA} calls. \begin{subentry} \iteme{= 0 :} no \ttt{LUEXEC} calls, i.e. only matrix-element and/or parton-shower treatment. \iteme{= 1 :} \ttt{LUEXEC} calls are made to generate fragmentation and decay chain. \iteme{= -1 :} no \ttt{LUEXEC} calls and no collapse of small jet systems into one or two particles (in \ttt{LUPREP}). \end{subentry} \iteme{MSTJ(106) :} (D=1) angular orientation in \ttt{LUEEVT} and \ttt{LUONIA}. \begin{subentry} \iteme{= 0 :} standard orientation of events, i.e. $\mathrm{q}$ along $+z$ axis and $\overline{\mathrm{q}}$ along $-z$ axis or in $xz$ plane with $p_x > 0$ for continuum events, and $\mathrm{g}_1\mathrm{g}_2\mathrm{g}_3$ or $\gamma\mathrm{g}_2\mathrm{g}_3$ in $xz$ plane with $\mathrm{g}_1$ or $\gamma$ along the $+z$ axis for onium events. \iteme{= 1 :} random orientation according to matrix elements. \end{subentry} \iteme{MSTJ(107) :} (D=0) radiative corrections to continuum events. \begin{subentry} \iteme{= 0 :} no radiative corrections. \iteme{= 1 :} initial-state radiative corrections (including weak effects for \ttt{MSTJ(102)=} 2 or 3). \end{subentry} \iteme{MSTJ(108) :} (D=2) calculation of $\alpha_{\mathrm{s}}$ for matrix-element alternatives. The \ttt{MSTU(111)} and \ttt{PARU(112)} values are automatically overwritten in \ttt{LUEEVT} or \ttt{LUONIA} calls accordingly. \begin{subentry} \iteme{= 0 :} fixed $\alpha_{\mathrm{s}}$ value as given in \ttt{PARU(111)}. \iteme{= 1 :} first-order formula is always used, with $\Lambda_{\mrm{QCD}}$ given by \ttt{PARJ(121)}. \iteme{= 2 :} first- or second-order formula is used, depending on value of \ttt{MSTJ(101)}, with $\Lambda_{\mrm{QCD}}$ given by \ttt{PARJ(121)} or \ttt{PARJ(122)}. \end{subentry} \iteme{MSTJ(109) :} (D=0) gives a possibility to switch from QCD matrix elements to some alternative toy models. Is not relevant for shower evolution, \ttt{MSTJ(101)=5}, where one can use \ttt{MSTJ(49)} instead. \begin{subentry} \iteme{= 0 :} standard QCD scenario. \iteme{= 1 :} a scalar gluon model. Since no second-order corrections are available in this scenario, one can only use this with \ttt{MSTJ(101) = 1} or \ttt{-1}. Also note that the event-as-a-whole angular distribution is for photon exchange only (i.e. no weak effects), and that no higher-order corrections to the total cross section are included. \iteme{= 2 :} an Abelian vector gluon theory, with the colour factors $C_F = 1$ ($= 4/3$ in QCD), $N_C = 0$ ($= 3$ in QCD) and $T_R = 3 n_f$ ($= n_f/2$ in QCD). If one selects $\alpha_{\mrm{Abelian}} = (4/3) \alpha_{\mrm{QCD}}$, the 3-jet cross section will agree with the QCD one, and differences are to be found only in 4-jets. The \ttt{MSTJ(109)=2} option has to be run with \ttt{MSTJ(110)=1} and \ttt{MSTJ(111)=0}; if need be, the latter variables will be overwritten by the program. \\ {\bf Warning:} second-order corrections give a large negative contribution to the 3-jet cross section, so large that the whole scenario is of doubtful use. In order to make the second-order options work at all, the 3-jet cross section is here by hand set exactly equal to zero for \ttt{MSTJ(101)=2}. It is here probably better to use the option \ttt{MSTJ(101)=3}, although this is not a consistent procedure either. \end{subentry} \iteme{MSTJ(110) :} (D=2) choice of second-order contributions to the 3-jet rate. \begin{subentry} \iteme{= 1 :} the GKS second-order matrix elements, i.e. the old {\tsc{Jetset}} standard. \iteme{= 2 :} the Zhu parametrization of the ERT matrix elements, based on the program of Kunszt and Ali, i.e. in historical sequence ERT/Kunszt/Ali/Zhu. The parametrization is available for $y =$ 0.01, 0.02, 0.03, 0.04 and 0.05. Values outside this range are put at the nearest border, while those inside it are given by a linear interpolation between the two nearest points. Since this procedure is rather primitive, one should try to work at one of the values given above. Note that no Abelian QCD parametrization is available for this option. \end{subentry} \iteme{MSTJ(111) :} (D=0) use of optimized perturbation theory for second-order matrix elements (it can also be used for first-order matrix elements, but here it only corresponds to a trivial rescaling of the $\alpha_{\mathrm{s}}$ argument). \begin{subentry} \iteme{= 0 :} no optimization procedure; i.e. $Q^2 = E_{\mrm{cm}}^2$. \iteme{= 1 :} an optimized $Q^2$ scale is chosen as $Q^2 = f E_{\mrm{cm}}^2$, where $f =$\ttt{PARJ(128)} for the total cross section $R$ factor, while $f =$\ttt{PARJ(129)} for the 3- and 4-jet rates. This $f$ value enters via the $\alpha_{\mathrm{s}}$, and also via a term proportional to $\alpha_{\mathrm{s}}^2 \ln f$. Some constraints are imposed; thus the optimized `3-jet' contribution to $R$ is assumed to be positive (for \ttt{PARJ(128)}), the total 3-jet rate is not allowed to be negative (for \ttt{PARJ(129)}), etc. However, there is no guarantee that the differential 3-jet cross section is not negative (and truncated to 0) somewhere (this can also happen with $f = 1$, but is then less frequent). The actually obtained $f$ values are stored in \ttt{PARJ(168)} and \ttt{PARJ(169)}, respectively. If an optimized $Q^2$ scale is used, then the $\Lambda_{\mrm{QCD}}$ (and $\alpha_{\mathrm{s}}$) should also be changed. With the value $f = 0.002$, it has been shown \cite{Bet89} that a $\Lambda_{\mrm{QCD}} = 0.100$ GeV gives a reasonable agreement; the parameter to be changed is \ttt{PARJ(122)} for a second-order running $\alpha_{\mathrm{s}}$. Note that, since the optimized $Q^2$ scale is sometimes below the charm threshold, the effective number of flavours used in $\alpha_{\mathrm{s}}$ may well be 4 only. If one feels that it is still appropriate to use 5 flavours (one choice might be as good as the other), it is necessary to put \ttt{MSTU(113)=5}. \end{subentry} \iteme{MSTJ(115) :} (D=1) documentation of continuum or onium events, in increasing order of completeness. \begin{subentry} \iteme{= 0 :} only the parton shower, the fragmenting partons and the generated hadronic system are stored in the \ttt{LUJETS} common block. \iteme{= 1 :} also a radiative photon is stored (for continuum events). \iteme{= 2 :} also the original $\e^+\e^-$ are stored (with \ttt{K(I,1)=21}). \iteme{= 3 :} also the $\gamma$ or $\gamma^* / \Z^0$ exchanged for continuum events, the onium state for resonance events is stored (with \ttt{K(I,1)=21}). \end{subentry} \iteme{MSTJ(116) :} (D=1) initialization of total cross section and radiative photon spectrum in \ttt{LUEEVT} calls. \begin{subentry} \iteme{= 0 :} never; cannot be used together with radiative corrections. \iteme{= 1 :} calculated at first call and then whenever \ttt{KFL} or \ttt{MSTJ(102)} is changed or \ttt{ECM} is changed by more than \ttt{PARJ(139)}. \iteme{= 2 :} calculated at each call. \iteme{= 3 :} everything is reinitialized in the next call, but \ttt{MSTJ(116)} is afterwards automatically put \ttt{=1} for use in subsequent calls. \end{subentry} \iteme{MSTJ(119) :} (I) check on need to reinitialize \ttt{LUXTOT}. \iteme{MSTJ(120) :} (R) type of continuum event generated with the matrix-element option (with the shower one, the result is always \ttt{=1}). \begin{subentry} \iteme{= 1 :} $\mathrm{q}\overline{\mathrm{q}}$. \iteme{= 2 :} $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}$. \iteme{= 3 :} $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g$ from Abelian (QED-like) graphs in matrix element. \iteme{= 4 :} $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g$ from non-Abelian (i.e. containing triple-gluon coupling) graphs in matrix element. \iteme{= 5 :} $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$. \end{subentry} \iteme{MSTJ(121) :} (R) flag set if a negative differential cross section was encountered in the latest \ttt{LUX3JT} call. Events are still generated, but maybe not quite according to the distribution one would like (the rate is set to zero in the regions of negative cross section, and the differential rate in the regions of positive cross section is rescaled to give the `correct' total 3-jet rate). \boxsep \iteme{PARJ(121) :}\label{p:PARJ121} (D=1.0 GeV) $\Lambda$ value used in first-order calculation of $\alpha_{\mathrm{s}}$ in the matrix-element alternative. \iteme{PARJ(122) :} (D=0.25 GeV) $\Lambda$ values used in second-order calculation of $\alpha_{\mathrm{s}}$ in the matrix-element alternative. \iteme{PARJ(123) :} (D=91.187 GeV) mass of $\mathrm{Z}^0$ as used in propagators for the QFD case. \iteme{PARJ(124) :} (D=2.489 GeV) width of $\mathrm{Z}^0$ as used in propagators for the QFD case. Overwritten at initialization if \ttt{MSTJ(102)=3}. \iteme{PARJ(125) :} (D=0.01) $y_{\mrm{cut}}$, minimum squared scaled invariant mass of any two partons in 3- or 4-jet events; the main user-controlled matrix-element cut. \ttt{PARJ(126)} provides an additional constraint. For each new event, it is additionally checked that the total 3- plus 4-jet fraction does not exceed unity; if so the effective $y$ cut will be dynamically increased. The actual $y$-cut value is stored in \ttt{PARJ(150)}, event by event. \iteme{PARJ(126) :} (D=2. GeV) minimum invariant mass of any two partons in 3- or 4-jet events; a cut in addition to the one above, mainly for the case of a radiative photon lowering the hadronic c.m. energy significantly. \iteme{PARJ(127) :} (D=1. GeV) is used as a safety margin for small colour-singlet jet systems, cf. \ttt{PARJ(32)}, specifically $\mathrm{q}\overline{\mathrm{q}}'$ masses in $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ 4-jet events and $\mathrm{g}\g$ mass in onium $\gamma\mathrm{g}\g$ events. \iteme{PARJ(128) :} (D=0.25) optimized $Q^2$ scale for the QCD $R$ (total rate) factor for the \ttt{MSTJ(111)=1} option is given by $Q^2 = f E_{\mrm{cm}}^2$, where $f =$\ttt{PARJ(128)}. For various reasons the actually used $f$ value may be increased compared with the nominal one; while \ttt{PARJ(128)} gives the nominal value, \ttt{PARJ(168)} gives the actual one for the current event. \iteme{PARJ(129) :} (D=0.002) optimized $Q^2$ scale for the 3- and 4-jet rate for the \ttt{MSTJ(111)=1} option is given by $Q^2 = f E_{\mrm{cm}}^2$, where $f =$\ttt{PARJ(129)}. For various reasons the actually used $f$ value may be increased compared with the nominal one; while \ttt{PARJ(129)} gives the nominal value, \ttt{PARJ(169)} gives the actual one for the current event. The default value is in agreement with the studies of Bethke \cite{Bet89}. \iteme{PARJ(131), PARJ(132) :} (D=2*0.) longitudinal polarizations $P_{\mrm{L}}^+$ and $P_{\mrm{L}}^-$ of incoming $\mathrm{e}^+$ and $\mathrm{e}^-$. \iteme{PARJ(133) :} (D=0.) transverse polarization $P_{\mrm{T}} = \sqrt{P_{\mrm{T}}^+ P_{\mrm{T}}^-}$, with $P_{\mrm{T}}^+$ and $P_{\mrm{T}}^-$ transverse polarizations of incoming $\mathrm{e}^+$ and $\mathrm{e}^-$. \iteme{PARJ(134) :} (D=0.) mean of transverse polarization directions of incoming $\mathrm{e}^+$ and $\mathrm{e}^-$, $\Delta \varphi = (\varphi^+ + \varphi^-) /2$, with $\varphi$ the azimuthal angle of polarization, leading to a shift in the $\varphi$ distribution of jets by $\Delta \varphi$. \iteme{PARJ(135) :} (D=0.01) minimum photon energy fraction (of beam energy) in initial-state radiation; should normally never be changed (if lowered too much, the fraction of events containing a radiative photon will exceed unity, leading to problems). \iteme{PARJ(136) :} (D=0.99) maximum photon energy fraction (of beam energy) in initial-state radiation; may be changed to reflect actual trigger conditions of a detector (but must always be larger than \ttt{PARJ(135)}). \iteme{PARJ(139) :} (D=0.2 GeV) maximum deviation of $E_{\mrm{cm}}$ from the corresponding value at last \ttt{LUXTOT} call, above which a new call is made if \ttt{MSTJ(116)=1}. \iteme{PARJ(141) :} (R) value of $R$, the ratio of continuum cross section to the lowest-order muon pair production cross section, as given in massless QED (i.e. three times the sum of active quark squared charges, possibly modified for polarization). \iteme{PARJ(142) :} (R) value of $R$ including quark-mass effects (for \ttt{MSTJ(102)=1}) and/or weak propagator effects (for \ttt{MSTJ(102)=2}). \iteme{PARJ(143) :} (R) value of $R$ as \ttt{PARJ(142)}, but including QCD corrections as given by \ttt{MSTJ(101)}. \iteme{PARJ(144) :} (R) value of $R$ as \ttt{PARJ(143)}, but additionally including corrections from initial-state photon radiation (if \ttt{MSTJ(107)=1}). Since the effects of heavy flavour thresholds are not simply integrable, the initial value of \ttt{PARJ(144)} is updated during the course of the run to improve accuracy. \iteme{PARJ(145) - PARJ(148) :} (R) absolute cross sections in nb as for the cases \ttt{PARJ(141) - PARJ(144)} above. \iteme{PARJ(150) :} (R) current effective matrix element cut-off $y_{\mrm{cut}}$, as given by \ttt{PARJ(125), PARJ(126)} and the requirements of having non-negative cross sections for 2-, 3- and 4-jet events. Not used in parton showers. \iteme{PARJ(151) :} (R) value of c.m. energy \ttt{ECM} at last \ttt{LUXTOT} call. \iteme{PARJ(152) :} (R) current first-order contribution to the 3-jet fraction; modified by mass effects. Not used in parton showers. \iteme{PARJ(153) :} (R) current second-order contribution to the 3-jet fraction; modified by mass effects. Not used in parton showers. \iteme{PARJ(154) :} (R) current second-order contribution to the 4-jet fraction; modified by mass effects. Not used in parton showers. \iteme{PARJ(155) :} (R) current fraction of 4-jet rate attributable to $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ events rather than $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g$ ones; modified by mass effects. Not used in parton showers. \iteme{PARJ(156) :} (R) has two functions when using second-order QCD. For a 3-jet event, it gives the ratio of the second-order to the total 3-jet cross section in the given kinematical point. For a 4-jet event, it gives the ratio of the modified 4-jet cross section, obtained when neglecting interference terms whose colour flow is not well defined, to the full unmodified one, all evaluated in the given kinematical point. Not used in parton showers. \iteme{PARJ(157) - PARJ(159) :} (I) used for cross-section calculations to include mass threshold effects to radiative photon cross section. What is stored is basic cross section, number of events generated and number that passed cuts. \iteme{PARJ(160) :} (R) nominal fraction of events that should contain a radiative photon. \iteme{PARJ(161) - PARJ(164) :} (I) give shape of radiative photon spectrum including weak effects. \iteme{PARJ(168) :} (R) actual $f$ value of current event in optimized perturbation theory for $R$; see \ttt{MSTJ(111)} and \ttt{PARJ(128)}. \iteme{PARJ(169) :} (R) actual $f$ value of current event in optimized perturbation theory for 3- and 4-jet rate; see \ttt{MSTJ(111)} and \ttt{PARJ(129)}. \iteme{PARJ(171) :} (R) fraction of cross section corresponding to the axial coupling of quark pair to the intermediate $\gamma^* / \Z^0$ state; needed for the Abelian gluon model 3-jet matrix element. \end{entry} \subsection{Examples} An ordinary $\e^+\e^-$ annihilation event in the continuum, at a c.m. energy of 40 GeV, may be generated with \begin{verbatim} CALL LUEEVT(0,40.) \end{verbatim} In this case a $\mathrm{q}\overline{\mathrm{q}}$ event is generated, including weak effects, followed by parton-shower evolution and fragmentation/decay treatment. Before a call to \ttt{LUEEVT}, however, a number of default values may be changed, e.g. \ttt{MSTJ(101)=2} to use second-order QCD matrix elements, giving a mixture of $\mathrm{q}\overline{\mathrm{q}}$, $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}$, $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}\g$, and $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ events, \ttt{MSTJ(102)=1} to have QED only, \ttt{MSTJ(104)=6} to allow $\t\overline{\mathrm{t}}$ production as well, \ttt{MSTJ(107)=1} to include initial-state photon radiation (including a treatment of the $\mathrm{Z}^0$ pole), \ttt{PARJ(123)=92.0} to change the $\mathrm{Z}^0$ mass, \ttt{PARJ(81)=0.3} to change the parton-shower $\Lambda$ value, or \ttt{PARJ(82)=1.5} to change the parton-shower cut-off. If initial-state photon radiation is used, some restrictions apply to how one can alternate the generation of events at different energies or with different $\mathrm{Z}^0$ mass, etc. These restrictions are not there for efficiency reasons (the extra time for recalculating the extra constants every time is small), but because it ties in with the cross-section calculations (see \ttt{PARJ(144)}). Most parameters can be changed independently of each other. However, if just one or a few parameters/switches are changed, one should not be surprised to find a rather bad agreement with the data, like e.g. a too low or high average hadron multiplicity. It is therefore usually necessary to retune one parameter related to the perturbative QCD description, like $\alpha_{\mathrm{s}}$ or $\Lambda$, one of the two parameters $a$ and $b$ of the Lund symmetric fragmentation function (since they are so strongly correlated, it is often not necessary to retune both of them), and the average fragmentation transverse momentum --- see Note~2 of the \ttt{MSTJ(101)} description for an example. For very detailed studies it may be necessary to retune even more parameters. The three-gluon and gluon--gluon--photon decays of $\Upsilon$ may be simulated by a call \begin{verbatim} CALL LUONIA(5,9.46) \end{verbatim} Unfortunately, with present top-mass limits, this routine will not be of much interest for toponium studies (weak decays will dominate). A typical program for analysis of $\e^+\e^-$ annihilation events at 100 GeV might look something like \begin{verbatim} COMMON/LUJETS/N,K(4000,5),P(4000,5),V(4000,5) COMMON/LUDAT1/MSTU(200),PARU(200),MSTJ(200),PARJ(200) COMMON/LUDAT2/KCHG(500,3),PMAS(500,4),PARF(2000),VCKM(4,4) COMMON/LUDAT3/MDCY(500,3),MDME(2000,2),BRAT(2000),KFDP(2000,5) MDCY(LUCOMP(111),1)=0 ! put pi0 stable MSTJ(107)=1 ! include initial-state radiation PARU(41)=1. ! use linear sphericity ..... ! other desired changes ..... ! initialize analysis statistics DO 100 IEVENT=1,1000 ! loop over events CALL LUEEVT(0,100.) ! generate new event IF(IEVENT.EQ.1) CALL LULIST(2) ! list first event CALL LUTABU(11) ! save particle composition ! statistics CALL LUEDIT(2) ! remove decayed particles CALL LUSPHE(SPH,APL) ! linear sphericity analysis IF(SPH.LT.0.) GOTO 100 ! too few particles in event for ! LUSPHE to work on it (unusual) CALL LUEDIT(31) ! orient event along axes above IF(IEVENT.EQ.1) CALL LULIST(2) ! list first treated event ..... ! fill analysis statistics CALL LUTHRU(THR,OBL) ! now do thrust analysis ..... ! more analysis statistics 100 CONTINUE ! CALL LUTABU(12) ! print particle composition ! statistics ..... ! print analysis statistics END \end{verbatim} \clearpage \section{Process Generation in PYTHIA} \label{s:PYTprocgen} Much can be said about the hard processes in {\tsc{Pythia}} and the way they are generated. Therefore the material has been split into three sections. In the current one the philo\-sophy underlying the event generation scheme is presented. Here we provide a generic description, where some special cases are swept under the carpet. In the next section, the existing processes are enumerated, with some comments about applications and limitations. Finally, in the third section the generation routines and common block switches are described. The section starts with a survey of parton distributions, followed by a detailed description of the simple $2 \to 2$ and $2 \to 1$ hard subprocess generation schemes, including pairs of resonances. This is followed by a few comments on more complicated configurations. \subsection{Parton Distributions} \label{ss:structfun} The parton-distribution function $f_i^a(x,Q^2)$ parametrizes the probability to find a parton $i$ with a fraction $x$ of the beam energy when the beam particle $a$ is probed by a hard scattering at virtuality scale $Q^2$. Usually the momentum-weighted combination $x f_i^a(x,Q^2)$ is used, for which the normalization condition $\sum_i \int_0^1 dx \, x f_i^a(x,Q^2) \equiv 1$ normally applies. The $Q^2$ dependence of parton distributions is perturbatively calculable, see section \ref{sss:initshowstruc}. The parton distributions in {\tsc{Pythia}} come in many shapes, as shown in the following. \subsubsection{Baryons} For protons, many sets exist on the market. These are obtained by fits to experimental data, constrained so that the $Q^2$ dependence is in accordance with the standard QCD evolution equations. The (new) default in \tsc{Pythia} is CTEQ2L \cite{Bot93}, a modern leading-order fit. Ten other sets are found in \tsc{Pythia}. The complete list is: \begin{Itemize} \item EHLQ sets 1 and 2 \cite{Eic84}; \item DO sets 1 and 2 \cite{Duk82}; \item the other CTEQ2 fits, namely CTEQ2M, CTEQ2MS, CTEQ2MF, CTEQ2ML, and CTEQ2D \cite{Bot93}; and \item the dynamically generated fit GRV LO (updated version) \cite{Glu92}. \end{Itemize} Of these, EHLQ, DO, CTEQ2L and GRV LO are leading-order parton distributions, while CTEQ2D are in the next-to-leading-order DIS scheme and the rest in the next-to-leading order $\br{\mrm{MS}}$ scheme. The EHLQ and DO sets are by now rather old, and are kept mainly for backwards compatibility. Since only Born-level matrix elements are included in the program, there is no particular reason to use higher-order parton distributions --- the resulting combination is anyway only good to leading-order accuracy. (Some higher-order corrections are effectively included by the parton-shower treatment, but there is no exact match.) There is a steady flow of new parton-distribution sets on the market. To keep track of all of them is a major work on its own. Therefore {\tsc{Pythia}} contains an interface to an external library of parton distribution functions, \tsc{Pdflib} \cite{Plo93}. This is a truly encyclopedic collection of almost all proton, pion and photon parton distributions proposed since the late 70's. Two dummy routines come with the {\tsc{Pythia}} package, so as to avoid problems with unresolved external references if \tsc{Pdflib} is not linked. One should also note that {\tsc{Pythia}} does not check the results, but assumes that sensible answers will be returned, also outside the nominal $(x, Q^2)$ range of a set. Only the sets that come with {\tsc{Pythia}} have been suitably modified to provide reasonable answers outside their nominal domain of validity. From the proton parton distributions, those of the neutron are obtained by isospin conjugation, i.e. $f_{\u}^{\mathrm{n}} = f_{\d}^{\mathrm{p}}$ and $f_{\d}^{\mathrm{n}} = f_{\u}^{\mathrm{p}}$. The program does allow for incoming beams of a number of hyperons: $\Lambda^0$, $\Sigma^{-,0,+}$, $\Xi^{-,0}$ and $\Omega^-$. Here one has essentially no experimental information. One could imagine to construct models in which valence $\mathrm{s}$ quarks are found at larger average $x$ values than valence $\u$ and $\d$ ones, because of the larger $\mathrm{s}$-quark mass. However, hyperon beams is a little-used part of the program, included only for a few specific studies. Therefore a simple approach has been taken, in which an average valence quark distribution is constructed as $f_{\mrm{val}} = (f_{\u,\mrm{val}}^{\mathrm{p}} + f_{\d,\mrm{val}}^{\mathrm{p}})/3$, according to which each valence quark in a hyperon is assumed to be distributed. Sea-quark and gluon distributions are taken as in the proton. Any proton parton distribution set may be used with this procedure. \subsubsection{Mesons and photons} Data on meson parton distributions are scarce, so only very few sets have been constructed, and only for the $\pi^{\pm}$. {\tsc{Pythia}} contains the Owens set 1 and 2 parton distributions \cite{Owe84}, which for a long time were essentially the only sets on the market, and the more recent dynamically generated GRV LO (updated version) \cite{Glu92a}. The first one is the default in {\tsc{Pythia}}. Further sets are found in \tsc{Pdflib} and can therefore be used by {\tsc{Pythia}}, just as described above for protons. Sets of photon parton distributions have been obtained as for hadrons; an additional complication comes from the necessity to handle the matching of the vector meson dominance (VMD) and the perturbative pieces in a consistent manner. New sets have been produced where this division is explicit and therefore especially well suited for applications to event generation\cite{Sch95}. The Schuler and Sj\"ostand set 1D is the default. Although the vector-meson philosophy is at the base, the details of the fits do not rely on pion data, but only on $F_2^{\gamma}$ data. Here follows a brief summary of relevant details. Photons obey a set of inhomogeneous evolution equations, where the inhomogeneous term is induced by $\gamma \to \mathrm{q}\overline{\mathrm{q}}$ branchings. The solution can be written as the sum of two terms, \begin{equation} f_a^{\gamma}(x,Q^2) = f_a^{\gamma,\mrm{NP}}(x,Q^2;Q_0^2) + f_a^{\gamma,\mrm{PT}}(x,Q^2;Q_0^2) ~, \end{equation} where the former term is a solution to the homogeneous evolution with a (non-perturbative) input at $Q=Q_0$ and the latter is a solution to the full inhomogeneous equation with boundary condition $f_a^{\gamma,\mrm{PT}}(x,Q_0^2;Q_0^2) \equiv 0$. One possible physics interpretation is to let $f_a^{\gamma,\mrm{NP}}$ correspond to $\gamma \leftrightarrow V$ fluctuations, where $V = \rho^0, \omega, \phi, \ldots$ is a set of vector mesons, and let $f_a^{\gamma,\mrm{PT}}$ correspond to perturbative (`anomalous') $\gamma \leftrightarrow \mathrm{q}\overline{\mathrm{q}}$ fluctuations. The discrete spectum of vector mesons can be combined with the continuous (in virtuality $k^2$) spectrum of $\mathrm{q}\overline{\mathrm{q}}$ fluctuations, to give \begin{equation} f_a^{\gamma}(x,Q^2) = \sum_V \frac{4\pi\alpha_{\mathrm{em}}}{f_V^2} f_a^{\gamma,V}(x,Q^2) + \frac{\alpha_{\mathrm{em}}}{2\pi} \, \sum_{\mathrm{q}} 2 e_{\mathrm{q}}^2 \, \int_{Q_0^2}^{Q^2} \frac{{\d} k^2}{k^2} \, f_a^{\gamma,\mathrm{q}\overline{\mathrm{q}}}(x,Q^2;k^2) ~, \end{equation} where each component $f^{\gamma,V}$ and $f^{\gamma,\mathrm{q}\overline{\mathrm{q}}}$ obeys a unit momentum sum rule. In sets 1 the $Q_0$ scale is picked at a low value, 0.6~GeV, where an identification of the non-perturbative component with a set of low-lying mesons appear natural, while sets 2 use a higher value, 2~GeV, where the validity of perturbation theory is better established. The data are not good enough to allow a precise determination of $\Lambda_{\mrm{QCD}}$. Therefore we use a fixed value $\Lambda^{(4)} = 200$~MeV, in agreement with conventional results for proton distributions. In the VMD component the $\rho^0$ and $\omega$ have been added coherently, so that $\u\overline{\mathrm{u}} : \d\overline{\mathrm{d}} = 4 : 1$ at $Q_0$. Unlike the $\mathrm{p}$, the $\gamma$ has a direct component where the photon acts as an unresolved probe. In the definition of $F_2^{\gamma}$ this adds a component $C^{\gamma}$, symbolically \begin{equation} F_2^{\gamma}(x,Q^2) = \sum_{\mathrm{q}} e_{\mathrm{q}}^2 \left[ f_{\mathrm{q}}^{\gamma} + f_{\overline{\mathrm{q}}}^{\gamma} \right] \otimes C_{\mathrm{q}} + f_{\mathrm{g}}^{\gamma} \otimes C_{\mathrm{g}} + C^{\gamma} ~. \end{equation} Since $C^{\gamma} \equiv 0$ in leading order, and since we stay with leading-order fits, it is permissible to neglect this complication. Numerically, however, it makes a non-negligible difference. We therefore make two kinds of fits, one DIS type with $C^{\gamma} = 0$ and one {$\overline{\mrm{MS}}$} type including the universal part of $C^{\gamma}$. When jet production is studied for real incoming photons, the standard evolution approach is reasonable also for heavy flavours, i.e. predominantly the $\c$, but with a lower cut-off $Q_0 \approx m_{\c}$ for $\gamma \to \c\overline{\mathrm{c}}$. Moving to deep inelastic scattering, $\mathrm{e}\gamma \to \mathrm{e} X$, there is an extra kinematical constraint: $W^2 = Q^2 (1-x)/x > 4 m_{\c}^2$. It is here better to use the `Bethe-Heitler' cross section for $\gamma^* \gamma \to \c\overline{\mathrm{c}}$. Therefore each distribution appears in two variants. For applications to real $\gamma$'s the parton distributions are calculated as the sum of a vector-meson part and an anomalous part including all five flavours. For applications to DIS, the sum runs over the same vector-meson part, an anomalous part and possibly a $C^{\gamma}$ part for the three light flavours, and a Bethe-Heitler part for $\c$ and $\b$. In addition to the SaS sets, {\tsc{Pythia}} also contains the Drees--Grassie set of parton distributions \cite{Dre85} and, as for the proton, there is an interface to the \tsc{Pdflib} library \cite{Plo93}. However, these sets do not allow a subdivision of the photon parton distributions into one VMD part and one anomalous part. This subdivision is necessary a sophisticated modelling of $\gamma\mathrm{p}$ and $\gamma\gamma$ events, see above and section \ref{sss:photoprod}. As an alternative, for the VMD part alone, the $\rho^0$ parton distribution can be found from the assumed equality \begin{equation} f^{\rho^0}_i = f^{\pi^0}_i = \frac{1}{2} \, (f^{\pi^+}_i + f^{\pi^-}_i) ~. \end{equation} Thus any $\pi^+$ parton distribution set, from any library, can be turned into a VMD $\rho^0$ set. The $\omega$ parton distribution is assumed the same, while the $\phi$ and $\mathrm{J}/\psi$ ones are handled in the very crude approximation $f^{\phi}_{\mathrm{s},\mrm{val}} = f^{\pi^+}_{\u,\mrm{val}}$ and $f^{\phi}_{\mrm{sea}} = f^{\pi^+}_{\mrm{sea}}$. therefore is default. The VMD part needs to be complemented by an anomalous part to make upp a full photon distribution. The latter is fully perturbatively calculable, given the lower cut-off scale $Q_0$. The SaS parametrization of the anomalous part is therefore used throughout for this purpose. The $Q_0$ scale can be set freely in the \ttt{PARP(15)} parameter. The $f_i^{\gamma,\mrm{anom}}$ distribution can be further decomposed, by the flavour and the $p_{\perp}$ of the original branching $\gamma \to \mathrm{q}\overline{\mathrm{q}}$. The flavour is distributed according to squared charge (plus flavour thresholds for heavy flavours) and the $p_{\perp}$ according to $\d p_{\perp}^2 / p_{\perp}^2$ in the range $Q_0 < p_{\perp} < Q$. At the branching scale, the photon only consists of a $\mathrm{q}\overline{\mathrm{q}}$ pair, with $x$ distribution $\propto x^2 + (1-x)^2$. A component $f_a^{\gamma,\mathrm{q}\overline{\mathrm{q}}}(x,Q^2;k^2)$, characterized by its $k \approx p_{\perp}$ and flavour, then is evolved homogeneously from $p_{\perp}$ to $Q$. For theoretical studies it is convenient to be able to access a specific component of this kind. Therefore also leading-order parametrizations of these decomposed distributions are available \cite{Sch95}. \subsubsection{Leptons} \label{sss:estructfun} Contrary to the hadron case, there is no necessity to introduce the parton-distribution function concept for leptons. A lepton can be considered as a point-like particle, with initial-state radiation handled by higher-order matrix elements. However, the parton distribution function approach offers a slightly simplified but very economical description of initial-state radiation effects for any hard process, also those for which higher-order corrections are not yet calculated. Parton distributions for electrons have been introduced in {\tsc{Pythia}}, but not yet for muons, i.e. currently $f_{\mu}^{\mu}(x, Q^2) = \delta (x-1)$. Also for the electron one is free to use a simple `unresolved' $\mathrm{e}$, $f_{\mathrm{e}}^{\mathrm{e}}(x, Q^2) = \delta(x-1)$, where the $\mathrm{e}$ retains the full original momentum. Electron parton distributions are calculable entirely from first principles, but different levels of approximation may be used. The parton-distribution formulae in {\tsc{Pythia}} are based on a next-to-leading-order exponentiated description, see ref. \cite{Kle89}, p. 34. The approximate behaviour is \begin{eqnarray} & & f_{\mathrm{e}}^{\mathrm{e}}(x,Q^2) \approx \frac{\beta}{2} (1-x)^{\beta/2-1} ~; \nonumber \\ & & \beta = \frac{2 \alpha_{\mathrm{em}}}{\pi} \left( \ln \frac{Q^2}{m_{\mathrm{e}}^2} -1 \right) ~. \end{eqnarray} The form is divergent but integrable for $x \to 1$, i.e. the electron likes to keep most of the energy. To handle the numerical precision problems for $x$ very close to unity, the parton distribution is set, by hand, to zero for $x > 0.999999$, and is rescaled upwards in the range $0.9999 < x < 0.999999$, in such a way that the total area under the parton distribution is preserved: \begin{equation} \left( f_{\mathrm{e}}^{\mathrm{e}}(x,Q^2) \right)_{\mrm{mod}} = \left\{ \begin{array}{ll} f_{\mathrm{e}}^{\mathrm{e}}(x,Q^2) & 0 \leq x \leq 0.9999 \\[2mm] \frac{\displaystyle 100^{\beta/2}}{\displaystyle 100^{\beta/2}-1} \, f_{\mathrm{e}}^{\mathrm{e}}(x,Q^2) & 0.9999 < x \leq 0.999999 \\[4mm] 0 & x > 0.999999 \, ~. \end{array} \right. \end{equation} The branchings $\mathrm{e} \to \mathrm{e} \gamma$, which are responsible for the softening of the $f_{\mathrm{e}}^{\mathrm{e}}$ parton distribution, also gives rise to a flow of photons. In photon-induced hard processes, the $f_{\gamma}^{\mathrm{e}}$ parton distribution can be used to describe the equivalent flow of photons. The formula used in the program is the simple first-order expression. There is some ambiguity in the choice of $Q^2$ range over which emissions should be included. The na\"{\i}ve (default) choice is \begin{equation} f_{\gamma}^{\mathrm{e}}(x,Q^2) = \frac{\alpha_{\mathrm{em}}}{2 \pi} \, \frac{1+(1-x)^2}{x} \, \ln \left( \frac{Q^2}{m_{\mathrm{e}}^2} \right) ~. \end{equation} Here it is assumed that only one scale enters the problem, namely that of the hard interaction, and that the scale of the branching $\mathrm{e} \to \mathrm{e} \gamma$ is bounded from above by the hard interaction scale. For a pure QCD or pure QED shower this is an appropriate procedure, cf. section \ref{sss:showermatching}, but in other cases it may not be optimal. In particular, for photoproduction the alternative that is probably most appropriate is \cite{Ali88}: \begin{equation} f_{\gamma}^{\mathrm{e}}(x,Q^2) = \frac{\alpha_{\mathrm{em}}}{2 \pi} \, \frac{1+(1-x)^2}{x} \, \ln \left( \frac{Q^2_{\mathrm{max}} (1-x)}{m_{\mathrm{e}}^2 \, x^2} \right) ~. \end{equation} Here $Q^2_{\mathrm{max}}$ is a user-defined cut for the range of scattered electron kinematics that is counted as photoproduction. Note that we now deal with two different $Q^2$ scales, one related to the hard subprocess itself, which appears as the argument of the parton distribution, and the other related to the scattering of the electron, which is reflected in $Q^2_{\mathrm{max}}$. In resolved photoproduction or resolved $\gamma\gamma$ interactions, one has to include the parton distributions for quarks and gluons inside the photon inside the electron. There are no published sets where results are directly presented in terms of quark and gluon distributions inside the electron. In the program, the $f_{\mathrm{q},\mathrm{g}}^{\mathrm{e}}$ are therefore obtained by a numerical convolution according to \begin{equation} f_{\mathrm{q},\mathrm{g}}^{\mathrm{e}}(x, Q^2) = \int_x^1 \frac{dx_{\gamma}}{x_{\gamma}} \, f_{\gamma}^{\mathrm{e}}(x_{\gamma}, Q^2) \, f^{\gamma}_{\mathrm{q},\mathrm{g}} \! \left( \frac{x}{x_{\gamma}}, Q^2 \right) ~, \label{pg:foldqgine} \end{equation} with $f^{\mathrm{e}}_{\gamma}$ as discussed above. The necessity for numerical convolution makes this parton distribution evaluation rather slow compared with the others; one should therefore only have it switched on for resolved photoproduction studies. One can obtain the positron distribution inside an electron, which is also the electron sea parton distribution, by a convolution of the two branchings $\mathrm{e} \to \mathrm{e} \gamma$ and $\gamma \to \e^+\e^-$; the result is \cite{Che75} \begin{equation} f_{\mathrm{e}^+}^{\mathrm{e}^-}(x,Q^2) = \frac{1}{2} \, \left\{ \frac{\alpha_{\mathrm{em}}}{2 \pi} \, \left( \ln \frac{Q^2}{m_{\mathrm{e}}^2} -1 \right) \right\}^2 \, \frac{1}{x} \, \left( \frac{4}{3} - x^2 - \frac{4}{3} x^3 + 2x(1+x) \ln x \right) ~. \end{equation} Finally, the program also contains the distribution of a transverse $\mathrm{W}^-$ inside an electron \begin{equation} f_{\mathrm{W}}^{\mathrm{e}}(x,Q^2) = \frac{\alpha_{\mathrm{em}}}{2 \pi} \, \frac{1}{4 \sin^2 \! \theta_W} \, \frac{1+(1-x)^2}{x} \, \ln \left( 1 + \frac{Q^2}{m_{\mathrm{W}}^2} \right) ~. \end{equation} \subsection{Kinematics and Cross section for a $2 \to 2$ Process} \label{ss:kinemtwo} In this section we begin the description of kinematics selection and cross-section calculation. The example is for the case of a $2 \to 2$ process, with final-state masses assumed to be vanishing. Later on we will expand to finite fixed masses, and to resonances. Consider two incoming beam particles in their c.m. frame, each with energy $E_{\mrm{beam}}$. The total squared c.m. energy is then $s = 4 E_{\mrm{beam}}^2$. The two partons that enter the hard interaction do not carry the total beam momentum, but only fractions $x_1$ and $x_2$, respectively, i.e. they have four-momenta \begin{eqnarray} p_1 & = & E_{\mrm{beam}}(x_1; 0, 0, x_1) ~, \nonumber \\ p_2 & = & E_{\mrm{beam}}(x_2; 0, 0, -x_2) ~. \end{eqnarray} There is no reason to put the incoming partons on the mass shell, i.e. to have time-like incoming four-vectors, since partons inside a particle are always virtual and thus space-like. These space-like virtualities are introduced as part of the initial-state parton-shower description, see section \ref{sss:initshowtrans}, but do not affect the formalism of this section. The one example where it would be appropriate to put a parton on the mass shell is for an incoming lepton beam, but even here the massless kinematics description is adequate as long as the c.m. energy is correctly calculated with masses. The squared invariant mass of the two partons is defined as \begin{equation} \hat{s} = (p_1 + p_2)^2 = x_1 \, x_2 \, s ~. \end{equation} Instead of $x_1$ and $x_2$, it is often customary to use $\tau$ and either $y$ or $x_{\mrm{F}}$: \begin{eqnarray} \tau & = & x_1 x_2 = \frac{\hat{s}}{s} ~; \\ y & = & \frac{1}{2} \ln \frac{x_1}{x_2} ~; \\ x_{\mrm{F}} & = & x_1 - x_2 ~. \end{eqnarray} In addition to $x_1$ and $x_2$, two additional variables are needed to describe the kinematics of a scattering $1 + 2 \to 3 + 4$. One corresponds to the azimuthal angle $\varphi$ of the scattering plane around the beam axis. This angle is always isotropically distributed for unpolarized incoming beam particles, and so need not be considered further. The other variable can be picked as $\hat{\theta}$, the polar angle of parton 3 in the c.m. frame of the hard scattering. The conventional choice is to use the variable \begin{equation} \hat{t} = (p_1 - p_3)^2 = (p_2 - p_4)^2 = - \frac{\hat{s}}{2} (1 - \cos \hat{\theta}) ~, \end{equation} with $\hat{\theta}$ defined as above. In the following, we will make use of both $\hat{t}$ and $\hat{\theta}$. It is also customary to define $\hat{u}$, \begin{equation} \hat{u} = (p_1 - p_4)^2 = (p_2 - p_3)^2 = - \frac{\hat{s}}{2} (1 + \cos \hat{\theta}) ~, \end{equation} but $\hat{u}$ is not an independent variable since \begin{equation} \hat{s} + \hat{t} + \hat{u} = 0 ~. \end{equation} If the two outgoing particles have masses $m_3$ and $m_4$, respectively, then the four-momenta in the c.m. frame of the hard interaction are given by \begin{equation} \hat{p}_{3,4} = \left( \frac{\hat{s} \pm (m_3^2-m_4^2)}{2\sqrt{\hat{s}}} , \pm \frac{\sqrt{\hat{s}}}{2} \, \beta_{34} \sin \hat{\theta}, 0, \pm \frac{\sqrt{\hat{s}}}{2} \, \beta_{34} \cos \hat{\theta} \right) ~, \end{equation} where \begin{equation} \beta_{34} = \sqrt{ \left( 1 - \frac{m_3^2}{\hat{s}} - \frac{m_4^2}{\hat{s}} \right)^2 - 4 \, \frac{m_3^2}{\hat{s}} \, \frac{m_4^2}{\hat{s}} } ~. \end{equation} Then $\hat{t}$ and $\hat{u}$ are modified to \begin{equation} \hat{t}, \hat{u} = - \frac{1}{2} \left\{ ( \hat{s} - m_3^2 - m_4^2 ) \mp \hat{s} \, \beta_{34} \cos \hat{\theta} \right\} ~, \end{equation} with \begin{equation} \hat{s} + \hat{t} + \hat{u} = m_3^2 + m_4^2 ~. \end{equation} The cross section for the process $1 + 2 \to 3 + 4$ may be written as \begin{eqnarray} \sigma & = & \int \int \int \d x_1 \, \d x_2 \, \d \hat{t} \, f_1(x_1, Q^2) \, f_2(x_2, Q^2) \, \frac{\d \hat{\sigma}}{\d \hat{t}} \nonumber \\ & = & \int \int \int \frac{\d \tau}{\tau} \, \d y \, \d \hat{t} \, x_1 f_1(x_1, Q^2) \, x_2 f_2(x_2, Q^2) \, \frac{\d \hat{\sigma}}{\d \hat{t}} ~. \label{pg:sigma} \end{eqnarray} The choice of $Q^2$ scale is ambiguous, and several alternatives are available in the program. For massless outgoing particles the default is the squared transverse momentum \begin{equation} Q^2 = \hat{p}_{\perp}^2 = \frac{\hat{s}}{4} \sin^2\hat{\theta} = \frac{\hat{t}\hat{u}}{\hat{s}} ~, \end{equation} which is modified to \begin{equation} Q^2 = \frac{1}{2} (m_{\perp 3}^2 + m_{\perp 4}^2) = \frac{1}{2} (m_3^2 + m_4^2) + \hat{p}_{\perp}^2 = \frac{1}{2} (m_3^2 + m_4^2) + \frac{\hat{t} \hat{u} - m_3^2 m_4^2}{\hat{s}} \end{equation} when masses are introduced. The mass term is selected such that, for $m_3 = m_4 = m$, the expression reduces to the squared transverse mass, $Q^2 = \hat{m}_{\perp}^2 = m^2 + \hat{p}_{\perp}^2$. The $\d \hat{\sigma}/\d \hat{t}$ expresses the differential cross section for a scattering, as a function of the kinematical quantities $\hat{s}$, $\hat{t}$ and $\hat{u}$. It is in this function that the physics of a given process resides. The performance of a machine is measured in terms of its luminosity $\cal L$, which is directly proportional to the number of particles in each bunch and to the bunch crossing frequency, and inversely proportional to the area of the bunches at the collision point. For a process with a $\sigma$ as given by eq.~(\ref{pg:sigma}), the differential event rate is given by $\sigma {\cal L}$, and the number of events collected over a given period of time \begin{equation} N = \sigma \, \int {\cal L} \, \d t ~. \end{equation} The program does not calculate the number of events, but only the integrated cross sections. \subsection{Resonance Production} \label{ss:kinemreson} The simplest way to produce a resonance is by a $2 \to 1$ process. If the decay of the resonance is not considered, the cross-section formula does not depend on $\hat{t}$, but takes the form \begin{equation} \sigma = \int \int \frac{\d \tau}{\tau} \, \d y \, x_1 f_1(x_1, Q^2) \, x_2 f_2(x_2, Q^2) \, \hat{\sigma}(\hat{s}) ~. \label{pg:sigmares} \end{equation} Here the physics is contained in the cross section $\hat{\sigma}(\hat{s})$. The $Q^2$ scale is usually taken to be $Q^2 = \hat{s}$. In published formulae, cross sections are often given in the zero-width approximation, i.e. $\hat{\sigma}(\hat{s}) \propto \delta (\hat{s} - m_R^2)$, where $m_R$ is the mass of the resonance. Introducing the scaled mass $\tau_R = m_R^2/s$, this corresponds to a delta function $\delta (\tau - \tau_R)$, which can be used to eliminate the integral over $\tau$. However, what we normally want to do is replace the $\delta$ function by the appropriate Breit--Wigner shape. For a resonance width $\Gamma_R$ this is achieved by the replacement \begin{equation} \delta (\tau - \tau_R) \to \frac{s}{\pi} \, \frac{m_R \Gamma_R}{(s \tau - m_R^2)^2 + m_R^2 \Gamma_R^2} ~. \label{pg:resshapeone} \end{equation} In this formula the resonance width $\Gamma_R$ is a constant. An improved description of resonance shapes is obtained if the width is made $\hat{s}$-dependent (occasionally also referred to as mass-dependent width, since $\hat{s}$ is not always the resonance mass), see e.g. \cite{Ber89}. To first approximation, this means that the expression $m_R \Gamma_R$ is to be replaced by $\hat{s} \Gamma_R / m_R$. To be more precise, in the program the quantity $H_R(\hat{s})$ is introduced, and the Breit--Wigner is written as \begin{equation} \delta (\tau - \tau_R) \to \frac{s}{\pi} \, \frac{H_R(s \tau)}{(s \tau - m_R^2)^2 + H_R^2(s \tau)} ~. \label{pg:resshapetwo} \end{equation} The $H_R$ factor is evaluated as a sum over all possible final-state channels, $H_R = \sum_f H_R^{(f)}$. Each decay channel may have its own $\hat{s}$ dependence, as follows. A decay to a fermion pair, $R \to \mathrm{f} \overline{\mathrm{f}}$, gives no contribution below threshold, i.e. for $\hat{s} < 4 m_{\mathrm{f}}^2$. Above threshold, $H_R^{(f)}$ is proportional to $\hat{s}$, multiplied by a threshold factor $\beta (3 - \beta^2)/2$ for the vector part of a spin 1 resonance, by $\beta^3$ for the axial vector part, and again by $\beta^3$ for a spin 0 resonance. Here $\beta = \sqrt{1 - 4m_{\mathrm{f}}^2/\hat{s}}$. For the decay into unequal masses, e.g. of the $\mathrm{W}^+$, corresponding but more complicated expressions are used. For decays into a quark pair, the universal first-order strong correction factor $1 + \alpha_{\mathrm{s}}(\hat{s}) / \pi$ is included in $H_R^{(f)}$. The second-order corrections are often known, but then are specific to each resonance, and are not included. An option exists for the $\gamma/\mathrm{Z}^0/\mathrm{Z}'^0$ resonances, where threshold effects due to $\mathrm{q}\overline{\mathrm{q}}$ bound-state formation are taken into account in a smeared-out, average sense, see eq.~(\ref{pp:threshenh}). For other decay channels, not into fermion pairs, the $\hat{s}$ dependence is typically more complicated. For instance, the decay $\H^0 \to \mathrm{W}^+ \mathrm{W}^-$ has a partial width proportional to $\hat{s}^2$, with a threshold factor $\beta^3$. Since a Higgs with $m_{\H} < 2 m_{\mathrm{W}}$ could still decay in this channel, it is in fact necessary to perform a two-dimensional integral over the $W^{\pm}$ Breit--Wigner mass distributions to obtain the correct result (and this has to be done numerically, at least in part). Fortunately, a Higgs particle lighter than $2 m_{\mathrm{W}}$ is sufficiently narrow that the integral only needs to be performed once and for all at initialization (whereas most other partial widths are recalculated whenever needed). Channels that proceed via loops, such as $\H \to \mathrm{g} \mathrm{g}$, also display complicated threshold behaviours. The coupling structure within the electroweak sector is usually (re)expressed in terms of gauge boson masses, $\alpha_{\mathrm{em}}$ and $\sin^2 \! \theta_W$, i.e. factors of $G_{\mathrm{F}}$ are replaced according to \begin{equation} \sqrt{2} G_{\mathrm{F}} = \frac{\pi \, \alpha_{\mathrm{em}}}{\sin^2 \! \theta_W \, m_{\mathrm{W}}^2} ~. \end{equation} Having done that, $\alpha_{\mathrm{em}}$ is allowed to run \cite{Kle89}, and is evaluated at the $\hat{s}$ scale. Thereby the relevant electroweak loop correction factors are recovered at the $m_{\mathrm{W}}/m_{\mathrm{Z}}$ scale. However, the option exists to go the other way and eliminate $\alpha_em$ in favour of $G_{\mathrm{F}}$. Currently $\sin^2 \! \theta_W$ is not allowed to run. For the Higgs particle, the couplings to fermions are proportional to the fermion masses; then also the masses are evaluated at the $\hat{s}$ scale. In summary, we see that an $\hat{s}$ dependence may enter several different ways into the $H_R^{(f)}$ expressions from which the total $H_R$ is built up. Also note that, with the exception of the term $(\mathrm{s} \tau - m_R^2)^2$ in the denominator of the Breit--Wigner, no memory remains of the nominal $m_R$ mass: everywhere else, what enters is the actual resonance mass $\sqrt{\hat{s}}$. When only decays to a specific final state $f$ are considered, the $H_R$ in the denominator remains the sum over all allowed decay channels, but the numerator only contains the $H_R^{(f)}$ term of the final state considered. If the combined production and decay process $i \to R \to f$ is considered, the same $\hat{s}$ dependence is implicit in the coupling structure of $i \to R$ as one would have had in $R \to i$, i.e. to first approximation there is a symmetry between couplings of a resonance to the initial and to the final state. The cross section $\hat{\sigma}$ is therefore, in the program, written in the form \begin{equation} \hat{\sigma}_{i \to R \to f}(\hat{s}) \propto \frac{\pi}{\hat{s}} \, \frac{H_R^{(i)}(\hat{s}) \, H_R^{(f)}(\hat{s})} {(\hat{s} - m_R^2)^2 + H_R^2(\hat{s})} ~. \label{pg:Hinoutsym} \end{equation} As a simple example, the cross section for the process $\mathrm{e}^- \br{\nu}_{\mathrm{e}} \to \mathrm{W}^- \to \mu^- \br{\nu}_{\mu}$ can be written as \begin{equation} \hat{\sigma}(\hat{s}) = 12 \, \frac{\pi}{\hat{s}} \, \frac{H_{\mathrm{W}}^{(i)}(\hat{s}) \, H_{\mathrm{W}}^{(f)}(\hat{s})} {(\hat{s} - m_{\mathrm{W}}^2)^2 + H_{\mathrm{W}}^2(\hat{s})} ~, \end{equation} where \begin{equation} H_{\mathrm{W}}^{(i)}(\hat{s}) = H_{\mathrm{W}}^{(f)}(\hat{s}) = \frac{\alpha_{\mathrm{em}}(\hat{s})}{24 \, \sin^2 \! \theta_W} \, \hat{s} ~. \end{equation} If the effects of several initial and/or final states are studied, it is straightforward to introduce an appropriate summation in the numerator. The analogy between the $H_R^{(f)}$ and $H_R^{(i)}$ cannot be pushed too far, however. The two differ in several important aspects. Firstly, colour factors appear reversed: the decay $R \to \mathrm{q}\overline{\mathrm{q}}$ contains a colour factor $N_C = 3$ enhancement, while $\mathrm{q}\overline{\mathrm{q}} \to R$ is instead suppressed by a factor $1/N_C = 1/3$. Secondly, the $1 + \alpha_{\mathrm{s}}(\hat{s}) / \pi$ first-order correction factor for the final state has to be replaced by a more complicated $K$ factor for the initial state. This factor is not usually known, or it is known (to first non-trivial order) but too lengthy to be included in the program. Thirdly, incoming partons as a rule are space-like. All the threshold suppression factors of the final state expressions are therefore irrelevant when production is considered. In sum, the $H_R^{(f)}$--$H_R^{(i)}$ analogy is mainly useful as a consistency cross-check, while the two usually are calculated separately. Exceptions include the rather messy loop structure involved in $\mathrm{g}\g \to \H^0$ and $\H^0 \to \mathrm{g}\g$, which is only coded once. It is of some interest to consider the observable resonance shape when the effects of parton distributions are included. In a hadron collider, to first approximation, parton distributions tend to have a behaviour roughly like $f(x) \propto 1/x$ for small $x$ --- this is why $f(x)$ is replaced by $xf(x)$ in eq. (\ref{pg:sigma}). Instead, the basic parton-distribution behaviour is shifted into the factor of $1/\tau$ in the integration phase space $\d \tau/\tau$, cf. eq.~(\ref{pg:sigmares}). When folded with the Breit--Wigner shape, two effects appear. One is that the overall resonance is tilted: the low-mass tail is enhanced and the high-mass one suppressed. The other is that an extremely long tail develops on the low-mass side of the resonance: when $\tau \to 0$, eq. (\ref{pg:Hinoutsym}) with $H_R(\hat{s}) \propto \hat{s}$ gives a $\hat{\sigma}(\hat{s}) \propto \hat{s} \propto \tau$, which exactly cancels the $1/\tau$ factor mentioned above. Na\"{\i}vely, the integral over $y$, $\int \d y = - \ln \tau$, therefore gives a net logarithmic divergence of the resonance shape when $\tau \to 0$. Clearly, it is then necessary to consider the shape of the parton distributions in more detail. At not-too-small $Q^2$, the evolution equations in fact lead to parton distributions more strongly peaked than $1/x$, typically with $xf(x) \propto x^{-0.3}$, and therefore a divergence like $\tau^{-0.3}$ in the cross-section expression. Eventually this divergence is regularized by a closing of the phase space, i.e. that $H_R(\hat{s})$ vanishes faster than $\hat{s}$, and by a less drastic small-$x$ parton-distribution behaviour when $Q^2 \approx \hat{s} \to 0$. The secondary peak at small $\tau$ may give a rather high cross section, which can even rival that of the ordinary peak around the nominal mass. This is the case, for instance, with $\mathrm{W}$ production. Such a peak has never been observed experimentally, but this is not surprising, since the background from other processes is overwhelming at low $\hat{s}$. Thus a lepton of one or a few GeV of transverse momentum is far more likely to come from the decay of a charm or bottom hadron than from a `$\mathrm{W}$' of a mass of a few GeV. When resonance production is studied, it is therefore important to set limits on the mass of the resonance, so as to agree with the experimental definition, at least to first approximation. If not, cross-section information given by the program may be very confusing. Another problem is that often the matrix elements really are valid only in the resonance region. The reason is that one usually includes only the simplest $s$-channel graph in the calculation. It is this `signal' graph that has a peak at the position of the resonance, where it (usually) gives much larger cross sections than the other `background' graphs. Away from the resonance position, `signal' and `background' may be of comparable order, or the `background' may even dominate. There is a quantum mechanical interference when some of the `signal' and `background' graphs have the same initial and final state, and this interference may be destructive or constructive. When the interference is non-negligible, it is no longer meaningful to speak of a `signal' cross section. As an example, consider the scattering of longitudinal $\mathrm{W}$'s, $\mathrm{W}^+_{\mrm{L}} \mathrm{W}^-_{\mrm{L}} \to \mathrm{W}^+_{\mrm{L}} \mathrm{W}^-_{\mrm{L}}$, where the `signal' process is $s$-channel exchange of a Higgs. This graph by itself is ill-behaved away from the resonance region. Destructive interference with `background' graphs such as $t$-channel exchange of a Higgs and $s$- and $t$-channel exchange of a $\gamma/\mathrm{Z}$ is required to save unitarity at large energies. In $\e^+\e^-$ colliders, the $f_{\mathrm{e}}^{\mathrm{e}}$ parton distribution is peaked at $x = 1$ rather than at $x = 0$. The situation therefore is the opposite, if one considers e.g. $\mathrm{Z}^0$ production in a machine running at energies above $m_{\mathrm{Z}}$: the tail towards lower masses is suppressed and the one towards higher masses enhanced, with a sharp secondary peak at around the nominal energy of the machine. Also in this case, an appropriate definition of cross sections therefore is necessary --- with additional complications due to the interference between $\gamma^*$ and $\mathrm{Z}^0$. When other processes are considered, problems of interference with background appears also here. Numerically the problems may be less pressing, however, since the secondary peak is occuring in a high-mass region, rather than in a more complicated low-mass one. Further, in $\e^+\e^-$ there is little uncertainty from the shape of the parton distributions. In $2 \to 2$ processes where a pair of resonances are produced, e.g. $\e^+\e^- \to \mathrm{Z}^0 \H^0$, cross section are almost always given in the zero-width approximation for the resonances. Here two substitutions of the type \begin{equation} 1 = \int \delta (m^2 - m_R^2) \, dm^2 \to \int \frac{1}{\pi} \, \frac{m_R \Gamma_R}{(m^2 - m_R^2)^2 + m_R^2 \Gamma_R^2} \, dm^2 \label{pg:resshapethree} \end{equation} are used to introduce mass distributions for the two resonance masses, i.e. $m_3^2$ and $m_4^2$. In the formula, $m_R$ is the nominal mass and $m$ the actually selected one. The phase-space integral over $x_1$, $x_1$ and $\hat{t}$ in eq. (\ref{pg:sigma}) is then extended to involve also $m_3^2$ and $m_4^2$. The effects of the mass-dependent width is only partly taken into account, by replacing the nominal masses $m_3^2$ and $m_4^2$ in the $\d \hat{\sigma} / \d \hat{t}$ expression by the actually generated ones (also e.g. in the relation between $\hat{t}$ and $\cos\hat{\theta}$), while the widths are evaluated at the nominal masses. This is the equivalent of a simple replacement of $m_R \Gamma_R$ by $\hat{s} \Gamma_R / m_R$ in the numerator of eq. (\ref{pg:resshapeone}), but not in the denominator. In addition, the full threshold dependence, i.e. the $\beta$-dependent factors, is not reproduced. There is no particular reason why the full mass-dependence could not be introduced, except for the extra work and time consumption needed for each process. In fact, the matrix elements for several $\gamma^* / \Z^0$ production processes do contain the full expressions. On the other hand, the matrix elements given in the literature are often valid only when the resonances are almost on the mass shell, since some graphs have been omitted. As an example, the process $\mathrm{q}\overline{\mathrm{q}} \to \mathrm{e}^- \br{\nu}_{\mathrm{e}} \mu^+ \nu_{\mu}$ is dominated by $\mathrm{q}\overline{\mathrm{q}} \to \mathrm{W}^- \mathrm{W}^+$ when each of the two lepton pairs is close to $m_{\mathrm{W}}$ in mass, but in general also receives contributions e.g. from $\mathrm{q}\overline{\mathrm{q}} \to \mathrm{Z}^0 \to \e^+\e^-$, followed by $\mathrm{e}^+ \to \br{\nu}_{\mathrm{e}} \mathrm{W}^+$ and $\mathrm{W}^+ \to \mu^+ \nu_{\mu}$. The latter contributions are neglected in cross sections given in the zero-width approximation. Processes with one final-state resonance and another ordinary final-state product, e.g. $\mathrm{q} \mathrm{g} \to \mathrm{W}^+ \mathrm{q}'$, are treated in the same spirit as the $2 \to 2$ processes with two resonances, except that only one mass need be selected according to a Breit--Wigner. \subsection{Cross-section Calculations} \label{ss:PYTcrosscalc} In the program, the variables used in the generation of a $2 \to 2$ process are $\tau$, $y$ and $z = \cos\hat{\theta}$. For a $2 \to 1$ process, the $z$ variable can be integrated out, and need therefore not be generated as part of the hard process, except when the allowed angular range of decays is restricted. In unresolved lepton beams, i.e. when $f_{\mathrm{e}}^{\mathrm{e}}(x) = \delta(x-1)$, the variables $\tau$ and/or $y$ may be integrated out. We will cover all these special cases towards the end of the section, and here concentrate on `standard' $2 \to 2$ and $2 \to 1$ processes. \subsubsection{The simple $2 \to 2$ processes} In the spirit of section \ref{ss:MCdistsel}, we want to select simple functions such that the true $\tau$, $y$ and $z$ dependence of the cross sections is approximately modelled. In particular, (almost) all conceivable kinematical peaks should be represented by separate terms in the approximate formulae. If this can be achieved, the ratio of the correct to the approximate cross sections will not fluctuate too much, but allow reasonable Monte Carlo efficiency. Therefore the variables are generated according to the distributions $h_{\tau}(\tau)$, $h_y(y)$ and $h_z(z)$, where normally \begin{eqnarray} h_{\tau}(\tau) & = & \frac{c_1}{{\cal I}_1} \, \frac{1}{\tau} + \frac{c_2}{{\cal I}_2} \, \frac{1}{\tau^2} + \frac{c_3}{{\cal I}_3} \, \frac{1}{\tau(\tau + \tau_R)} + \frac{c_4}{{\cal I}_4} \, \frac{1}{(s\tau - m_R^2)^2 + m_R^2 \Gamma_R^2} \nonumber \\ & & + \frac{c_5}{{\cal I}_5} \, \frac{1}{\tau(\tau + \tau_{R'})} + \frac{c_6}{{\cal I}_6} \, \frac{1}{(s\tau - m_{R'}^2)^2 + m_{R'}^2 \Gamma_{R'}^2} ~, \\[1mm] h_y(y) & = & \frac{c_1}{{\cal I}_1} \, (y - y_{\mathrm{min}}) + \frac{c_2}{{\cal I}_2} \, (y_{\mathrm{max}} - y) + \frac{c_3}{{\cal I}_3} \, \frac{1}{\cosh y} ~, \\[1mm] h_z(z) & = & \frac{c_1}{{\cal I}_1} + \frac{c_2}{{\cal I}_2} \, \frac{1}{a-z} + \frac{c_3}{{\cal I}_3} \, \frac{1}{a+z} + \frac{c_4}{{\cal I}_4} \, \frac{1}{(a-z)^2} + \frac{c_5}{{\cal I}_5} \, \frac{1}{(a+z)^2} ~. \label{pg:hforms} \end{eqnarray} Here each term is separately integrable, with an invertible primitive function, such that generation of $\tau$, $y$ and $z$ separately is a standard task, as described in section \ref{ss:MCdistsel}. In the following we describe the details of the scheme, including the meaning of the coefficients $c_i$ and ${\cal I}_i$, which are separate for $\tau$, $y$ and $z$. The first variable to be selected is $\tau$. The range of allowed values, $\tau_{\mathrm{min}} \leq \tau \leq \tau_{\mathrm{max}}$, is generally constrained by a number of user-defined requirements. A cut on the allowed mass range is directly reflected in $\tau$, a cut on the $p_{\perp}$ range indirectly. The first two terms of $h_{\tau}$ are intended to represent a smooth $\tau$ dependence, as generally obtained in processes which do not receive contributions from $s$-channel resonances. Also $s$-channel exchange of essentially massless particles ($\gamma$, $\mathrm{g}$, light quarks and leptons) are accounted for, since these do not produce any separate peaks at non-vanishing $\tau$. The last four terms of $h_{\tau}$ are there to catch the peaks in the cross section from resonance production. These terms are only included when needed. Each resonance is represented by two pieces, a first to cover the interference with graphs which peak at $\tau =0$, plus the variation of parton distributions, and a second to approximate the Breit--Wigner shape of the resonance itself. The subscripts $R$ and $R'$ denote values pertaining to the two resonances, with $\tau_R = m_R^2/s$. Currently there is only one process where the full structure with two resonances is used, namely $\mathrm{f} \overline{\mathrm{f}} \to \gamma^*/\mathrm{Z}^0/\mathrm{Z}'^0$. Otherwise either one or no resonance peak is taken into account. The kinematically allowed range of $y$ values is constrained by $\tau$, $|y| \leq - \frac{1}{2} \ln\tau$, and you may impose additional cuts. Therefore the allowed range $y_{\mathrm{min}} \leq y \leq y_{\mathrm{max}}$ is only constructed after $\tau$ has been selected. The first two terms of $h_y$ give a fairly flat $y$ dependence --- for processes which are symmetric in $y \leftrightarrow -y$, they will add to give a completely flat $y$ spectrum between the allowed limits. In principle, the natural subdivision would have been one term flat in $y$ and one forward--backward asymmetric, i.e. proportional to $y$. The latter is disallowed by the requirement of positivity, however. The $y - y_{\mathrm{min}}$ and $y_{\mathrm{max}} - y$ terms actually used give the same amount of freedom, but respect positivity. The third term is peaked at around $y = 0$, and represents the bias of parton distributions towards this region. The allowed $z = \cos\hat{\theta}$ range is na\"{\i}vely $-1 \leq z \leq 1$. However, most cross sections are divergent for $z \to \pm 1$, so some kind of regularization is necessary. Normally one requires $p_{\perp} \geq p_{\perp\mathrm{min}}$, which translates into $z^2 \leq 1 - 4 p_{\perp\mathrm{min}}^2/(\tau s)$ for massless outgoing particles. Since again the limits depend on $\tau$, the selection of $z$ is done after that of $\tau$. Additional requirements may constrain the range further. In particular, a $p_{\perp\mathrm{max}}$ constraint may split the allowed $z$ range into two, i.e. $z_{-\mathrm{min}} \leq z \leq z_{-\mathrm{max}}$ or $z_{+\mathrm{min}} \leq z \leq z_{+\mathrm{max}}$. An unsplit range is represented by $z_{-\mathrm{max}} = z_{+\mathrm{min}} = 0$. For massless outgoing particles the parameter $a = 1$ in $h_z$, such that the five terms represent a piece flat in angle and pieces peaked as $1/\hat{t}$, $1/\hat{u}$, $1/\hat{t}^2$, and $1/\hat{u}^2$, respectively. For non-vanishing masses one has $a = 1 + 2 m_3^2 m_4^2/\hat{s}^2$. In this case, the full range $-1 \leq z \leq 1$ is therefore available --- physically, the standard $\hat{t}$ and $\hat{u}$ singularities are regularized by the masses $m_3$ and $m_4$. For each of the terms, the ${\cal I}_i$ coefficients represent the integral over the quantity multiplying the coefficient $c_i$; thus, for instance: \begin{eqnarray} h_{\tau}: & & {\cal I}_1 = \int \frac{\d \tau}{\tau} = \ln \left( \frac{\tau_{\mathrm{max}}}{\tau_{\mathrm{min}}} \right) ~, \nonumber \\ & & {\cal I}_2 = \int \frac{\d \tau}{\tau^2} = \frac{1}{\tau_{\mathrm{min}}} - \frac{1}{\tau_{\mathrm{max}}} ~; \nonumber \\ h_y: & & {\cal I}_1 = \int (y - y_{\mathrm{min}}) \, \d y = \frac{1}{2} (y_{\mathrm{max}} - y_{\mathrm{min}})^2 ~; \nonumber \\ h_z: & & {\cal I}_1 = \int \d z = (z_{-\mathrm{max}} - z_{-\mathrm{min}}) + (z_{+\mathrm{max}} - z_{+\mathrm{min}}) , \nonumber \\ & & {\cal I}_2 = \int \frac{\d z}{a-z} = \ln \left( \frac{(a-z_{-\mathrm{min}})(a-z_{+\mathrm{min}})} {(a-z_{-\mathrm{max}})(a-z_{-\mathrm{min}})} \right) ~. \end{eqnarray} The $c_i$ coefficients are normalized to unit sum for $h_{\tau}$, $h_y$ and $h_z$ separately. They have a simple interpretation, as the probability for each of the terms to be used in the preliminary selection of $\tau$, $y$ and $z$, respectively. The variation of the cross section over the allowed phase space is explored in the initialization procedure of a {\tsc{Pythia}} run, and based on this knowledge the $c_i$ are optimized so as to give functions $h_{\tau}$, $h_y$ and $h_z$ that closely follow the general behaviour of the true cross section. For instance, the coefficient $c_4$ in $h_{\tau}$ is to be made larger the more the total cross section is dominated by the region around the resonance mass. The phase-space points tested at initialization are put on a grid, with the number of points in each dimension given by the number of terms in the respective $h$ expression, and with the position of each point given by the median value of the distribution of one of the terms. For instance, the $\d \tau / \tau$ distribution gives a median point at $\sqrt{\tau_{\mathrm{min}}\tau_{\mathrm{max}}}$, and $\d \tau / \tau^2$ has the median $2 \tau_{\mathrm{min}} \tau_{\mathrm{max}} / (\tau_{\mathrm{min}} + \tau_{\mathrm{max}})$. Since the allowed $y$ and $z$ ranges depend on the $\tau$ value selected, then so do the median points defined for these two variables. With only a limited set of phase-space points studied at the initialization, the `optimal' set of coefficients is not uniquely defined. To be on the safe side, 40\% of the total weight is therefore assigned evenly between all allowed $c_i$, whereas the remaining 60\% are assigned according to the relative importance surmised, under the constraint that no coefficient is allowed to receive a negative contribution from this second piece. After a preliminary choice has been made of $\tau$, $y$ and $z$, it is necessary to find the weight of the event, which is to be used to determine whether to keep it or generate another one. Using the relation $\d \hat{t} = \hat{s} \, \beta_{34} \, \d z / 2$, eq.~(\ref{pg:sigma}) may be rewritten as \begin{eqnarray} \sigma & = & \int \int \int \frac{\d \tau}{\tau} \, \d y \, \frac{\hat{s} \beta_{34}}{2} \d z \, x_1 f_1(x_1, Q^2) \, x_2 f_2(x_2, Q^2) \, \frac{\d \hat{\sigma}}{\d \hat{t}} \nonumber \\[1mm] & = & \frac{\pi}{s} \int h_{\tau}(\tau) \, \d \tau \int h_y(y) \, \d y \int h_z(z) \, \d z \, \beta_{34} \, \frac{x_1 f_1(x_1, Q^2) \, x_2 f_2(x_2, Q^2)} {\tau^2 h_{\tau}(\tau) \, h_y(y) \, 2 h_z(z)} \, \frac{\hat{s}^2}{\pi} \frac{\d \hat{\sigma}}{\d \hat{t}} \nonumber \\[1mm] & = & \left\langle \frac{\pi}{s} \, \frac{\beta_{34}}{\tau^2 h_{\tau}(\tau) \, h_y(y) \, 2 h_z(z)} \, x_1 f_1(x_1, Q^2) \, x_2 f_2(x_2, Q^2) \, \frac{\hat{s}^2}{\pi} \frac{\d \hat{\sigma}}{\d \hat{t}} \right\rangle ~. \label{pg:sigmamap} \end{eqnarray} In the middle line, a factor of $1 = h_{\tau}/h_{\tau}$ has been introduced to rewrite the $\tau$ integral in terms of a phase space of unit volume: $\int h_{\tau}(\tau) \, \d \tau = 1$ according to the relations above. Correspondingly for the $y$ and $z$ integrals. In addition, factors of $1 = \hat{s}/ (\tau s)$ and $1 = \pi / \pi$ are used to isolate the dimensionless cross section $(\hat{s}^2/\pi) \, \d \hat{\sigma} / \d \hat{t}$. The content of the last line is that, with $\tau$, $y$ and $z$ selected according to the expressions $h_{\tau}(\tau)$, $h_y(y)$ and $h_z(z)$, respectively, the cross section is obtained as the average of the final expression over all events. Since the $h$'s have been picked to give unit volume, there is no need to multiply by the total phase-space volume. As can be seen, the cross section for a given Monte Carlo event is given as the product of four factors, as follows: \begin{Enumerate} \item The $\pi/s$ factor, which is common to all events, gives the overall dimensions of the cross section, in GeV$^{-2}$. Since the final cross section is given in units of mb, the conversion factor of 1 GeV$^{-2} = 0.3894$ mb is also included here. \item Next comes the `Jacobian', which compensates for the change from the original to the final phase-space volume. \item The parton-distribution function weight is obtained by making use of the parton distribution libraries in {\tsc{Pythia}} or externally. The $x_1$ and $x_2$ values are obtained from $\tau$ and $y$ via the relations $x_{1,2} = \sqrt{\tau} \exp(\pm y)$. \item Finally, the dimensionless cross section $(\hat{s}^2/\pi) \, \d \hat{\sigma} / \d \hat{t}$ is the quantity that has to be coded for each process separately, and where the physics content is found. \end{Enumerate} Of course, the expression in the last line is not strictly necessary to obtain the cross section by Monte Carlo integration. One could also have used eq.~(\ref{pg:sigma}) directly, selecting phase-space points evenly in $\tau$, $y$ and $\hat{t}$, and averaging over those Monte Carlo weights. Clearly this would be much simpler, but the price to be paid is that the weights of individual events could fluctuate wildly. For instance, if the cross section contains a narrow resonance, the few phase-space points that are generated in the resonance region obtain large weights, while the rest do not. With our procedure, a resonance would be included in the $h_{\tau}(\tau)$ factor, so that more events would be generated at around the appropriate $\tau_R$ value (owing to the $h_{\tau}$ numerator in the phase-space expression), but with these events assigned a lower, more normal weight (owing to the factor $1/h_{\tau}$ in the weight expression). Since the weights fluctuate less, fewer phase-space points need be selected to get a reasonable cross-section estimate. In the program, the cross section is obtained as the average over all phase-space points generated. The events actually handed on to the user should have unit weight, however (an option with weighted events exists, but does not represent the mainstream usage). At initialization, after the $c_i$ coefficients have been determined, a search inside the allowed phase-space volume is therefore made to find the maximum of the weight expression in the last line of eq.~(\ref{pg:sigmamap}). In the subsequent generation of events, a selected phase-space point is then retained with a probability equal to the weight in the point divided by the maximum weight. Only the retained phase-space points are considered further, and generated as complete events. The search for the maximum is begun by evaluating the weight in the same grid of points as used to determine the $c_i$ coefficients. The point with highest weight is used as starting point for a search towards the maximum. In unfortunate cases, the convergence could be towards a local maximum which is not the global one. To somewhat reduce this risk, also the grid point with second-highest weight is used for another search. After initialization, when events are generated, a warning message will be given by default at any time a phase-space point is selected where the weight is larger than the maximum, and thereafter the maximum weight is adjusted to reflect the new knowledge. This means that events generated before this time have a somewhat erroneous distribution in phase space, but if the maximum violation is rather modest the effects should be negligible. The estimation of the cross section is not affected by any of these considerations, since the maximum weight does not enter into eq. (\ref{pg:sigmamap}). For $2 \to 2$ processes with identical final-state particles, the symmetrization factor of $1/2$ is explicitly included at the end of the $\d \hat{\sigma} / \d \hat{t}$ calculation. In the final cross section, a factor of 2 is retrieved because of integration over the full phase space (rather than only half of it). That way, no special provisions are needed in the phase-space integration machinery. \subsubsection{Resonance production} We have now covered the simple $2 \to 2$ case. In a $2 \to 1$ process, the $\hat{t}$ integral is absent, and the differential cross section $\d \hat{\sigma} / \d \hat{t}$ is replaced by $\hat{\sigma}(\hat{s})$. The cross section may now be written as \begin{eqnarray} \sigma & = & \int \int \frac{\d \tau}{\tau} \, \d y \, x_1 f_1(x_1, Q^2) \, x_2 f_2(x_2, Q^2) \, \hat{\sigma}(\hat{s}) \nonumber \\ & = & \frac{\pi}{s} \int h_{\tau}(\tau) \, \d \tau \int h_y(y) \, \d y \, \frac{x_1 f_1(x_1, Q^2) \, x_2 f_2(x_2, Q^2)} {\tau^2 h_{\tau}(\tau) \, h_y(y)} \, \frac{\hat{s}}{\pi} \hat{\sigma}(\hat{s}) \nonumber \\ & = & \left\langle \frac{\pi}{s} \, \frac{1}{\tau^2 h_{\tau}(\tau) \, h_y(y)} \, x_1 f_1(x_1, Q^2) \, x_2 f_2(x_2, Q^2) \, \frac{\hat{s}}{\pi} \hat{\sigma}(\hat{s}) \right\rangle ~. \label{pg:sigmamapres} \end{eqnarray} The structure is thus exactly the same, but the $z$-related pieces are absent, and the r\^ole of the dimensionless cross section is played by $(\hat{s}/\pi) \hat{\sigma}(\hat{s})$. If the range of allowed decay angles of the resonance is restricted, e.g. by requiring the decay products to have a minimum transverse momentum, effectively this translates into constraints on the $z = \cos\hat{\theta}$ variable of the $2 \to 2$ process. The difference is that the angular dependence of a resonance decay is trivial, and that therefore the $z$-dependent factor can be easily evaluated. For a spin-0 resonance, which decays isotropically, the relevant weight is simply $(z_{-\mathrm{max}} - z_{-\mathrm{min}})/2 + (z_{+\mathrm{max}} - z_{+\mathrm{min}})/2$. For a transversely polarized spin-1 resonance the expression is, instead, \begin{equation} \frac{3}{8}(z_{-\mathrm{max}} - z_{-\mathrm{min}}) + \frac{3}{8}(z_{+\mathrm{max}} - z_{+\mathrm{min}}) + \frac{1}{8}(z_{-\mathrm{max}} - z_{-\mathrm{min}})^3 + \frac{1}{8}(z_{+\mathrm{max}} - z_{+\mathrm{min}})^3 ~. \end{equation} Since the allowed $z$ range could depend on $\tau$ and/or $y$ (it does for a $p_{\perp}$ cut), the factor has to be evaluated for each individual phase-space point and included in the expression of eq. (\ref{pg:sigmamapres}). For $2 \to 2$ processes where either of the final-state particles is a resonance, or both, an additional choice has to be made for each resonance mass, eq.~(\ref{pg:resshapethree}). Since the allowed $\tau$, $y$ and $z$ ranges depend on $m_3^2$ and $m_4^2$, the selection of masses have to precede the choice of the other phase-space variables. Just as for the other variables, masses are not selected uniformly over the allowed range, but are rather distributed according to a function $h_m(m^2) \, dm^2$, with a compensating factor $1/h_m(m^2)$ in the `Jacobian'. The functional form picked is normally \begin{equation} h_m(m^2) = \frac{c_1}{{\cal I}_1} \, \frac{1}{\pi} \, \frac{m_R \Gamma_R}{(m^2 - m_R^2)^2 + m_R^2 \Gamma_R^2} + \frac{c_2}{{\cal I}_2} + \frac{c_3}{{\cal I}_3} \, \frac{1}{m^2} + \frac{c_4}{{\cal I}_4} \, \frac{1}{m^4} ~. \label{pg:reshdist} \end{equation} The definition of the ${\cal I}_i$ integrals is analogous to the one before. The $c_i$ coefficients are not found by optimization, but predetermined, normally to $c_1 = 0.8$, $c_2 = c_3 =0.1$, $c_4 = 0$. Clearly, had the phase space and the cross section been independent of $m_3^2$ and $m_4^2$, the optimal choice would have been to put $c_1 = 1$ and have all other $c_i$ vanishing --- then the $1/h_m$ factor of the `Jacobian' would exactly have cancelled the Breit--Wigner of eq.~(\ref{pg:resshapethree}) in the cross section. The second and the third terms are there to cover the possibility that the cross section does not die away quite as fast as given by the na\"{\i}ve Breit--Wigner shape. In particular, the third term covers the possibility of a secondary peak at small $m^2$, in a spirit slightly similar to the one discussed for resonance production in $2 \to 1$ processes. The fourth term is only used for the process $\mathrm{f} \overline{\mathrm{f}} \to (\gamma^* / \Z^0)(\gamma^* / \Z^0)$, where the $\gamma$ propagator guarantees that the cross section does have a significant secondary peak for $m^2 \to 0$. Therefore here the choice is $c_1 = 0.4$, $c_2 = 0.05$, $c_3 = 0.3$ and $c_4 = 0.25$. A few special tricks have been included to improve efficiency when the allowed mass range of resonances is constrained by kinematics or by user cuts. For instance, if a pair of equal or charge-conjugate resonances are produced, such as in $\e^+\e^- \to \mathrm{W}^+ \mathrm{W}^-$, use is made of the constraint that the lighter of the two has to have a mass smaller than half the c.m. energy. \subsubsection{Lepton beams} Lepton beams have to be handled slightly differently from what has been described so far. One also has to distinguish between a lepton for which parton distributions are included and one which is treated as an unresolved point-like particle. The necessary modifications are the same for $2 \to 2$ and $2 \to 1$ processes, however, since the $\hat{t}$ degree of freedom is unaffected. If one incoming beam is an unresolved lepton, the corresponding parton-distribution piece collapses to a $\delta$ function. This function can be used to integrate out the $y$ variable: $\delta(x_{1,2} - 1) = \delta(y \pm (1/2) \ln \tau)$. It is therefore only necessary to select the $\tau$ and the $z$ variables according to the proper distributions, with compensating weight factors, and only one set of parton distributions has to be evaluated explicitly. If both incoming beams are unresolved leptons, both the $\tau$ and the $y$ variables are trivially given: $\tau = 1$ and $y = 0$. Parton-distribution weights disappear completely. For a $2 \to 2$ process, only the $z$ selection remains to be performed, while a $2 \to 1$ process is completely specified, i.e. the cross section is a simple number that only depends on the c.m. energy. For a resolved electron, the $f_{\mathrm{e}}^{\mathrm{e}}$ parton distribution is strongly peaked towards $x = 1$. This affects both the $\tau$ and the $y$ distributions, which are not well described by either of the pieces in $h_{\tau}(\tau)$ or $h_y(y)$ in processes with interacting $\mathrm{e}^{\pm}$. (Processes which involve e.g. the $\gamma$ content of the $\mathrm{e}$ are still well simulated, since $f_{\gamma}^{\mathrm{e}}$ is peaked at small $x$.) If both parton distributions are peaked close to 1, the $h_{\tau}(\tau)$ expression in eq. (\ref{pg:hforms}) is therefore increased with one additional term of the form $h_{\tau}(\tau) \propto 1 / (1 - \tau)$, with coefficients $c_7$ and ${\cal I}_7$ determined as before. The divergence when $\tau \to 1$ is cut off by our regularization procedure for the $f_{\mathrm{e}}^{\mathrm{e}}$ parton distribution; therefore we only need consider $\tau < 1 - 2 \times 10^{-6}$. Correspondingly, the $h_y(y)$ expression is expanded with a term $1/(1 - \exp(y-y_0))$ when incoming beam number 1 consists of a resolved $\mathrm{e}^{\pm}$, and with a term $1/(1 - \exp(-y-y_0))$ when incoming beam number 2 consists of a resolved $\mathrm{e}^{\pm}$. Both terms are present for an $\e^+\e^-$ collider, only one for an $\e\p$ one. The coefficient $y_0 = - (1/2) \ln \tau$ is the na\"{\i}ve kinematical limit of the $y$ range, $|y| < y_0$. From the definitions of $y$ and $y_0$ it is easy to see that the two terms above correspond to $1/(1-x_1)$ and $1/(1-x_2)$, respectively, and thus are again regularized by our parton-distribution function cut-off. Therefore the integration ranges are $y < y_0 -10^{-6}$ for the first term and $y > - y_0 + 10^{-6}$ for the second one. \subsubsection{Mixing processes} \label{sss:mixingproc} In the cross-section formulae given so far, we have deliberately suppressed a summation over the allowed incoming flavours. For instance, the process $\mathrm{f}\overline{\mathrm{f}} \to \mathrm{Z}^0$ in a hadron collider receives contributions from $\u\overline{\mathrm{u}} \to \mathrm{Z}^0$, $\d\overline{\mathrm{d}} \to \mathrm{Z}^0$, $\mathrm{s}\overline{\mathrm{s}} \to \mathrm{Z}^0$, and so on. These contributions share the same basic form, but differ in the parton-distribution weights and (usually) in a few coupling constants in the hard matrix elements. It it therefore convenient to generate the terms together, as follows: \begin{Enumerate} \item A phase-space point is picked, and all common factors related to this choice are evaluated, i.e. the `Jacobian' and the common pieces of the matrix elements (e.g. for a $\mathrm{Z}^0$ the basic Breit--Wigner shape, excluding couplings to the initial flavour). \item The parton-distribution-function library is called to produce all the parton distributions, at the relevant $x$ and $Q^2$ values, for the two incoming beams. \item A loop is made over the two incoming flavours, one from each beam particle. For each allowed set of incoming flavours, the full matrix-element expression is put together, using the common pieces and the flavour-dependent couplings. This is multiplied by the common factors and the parton-distribution weights to obtain a cross-section weight. \item Each allowed flavour combination is stored as a separate entry in a table, together with its weight. In addition, a summed weight is calculated. \item The phase-space point is kept or rejected, according to a comparison of the summed weight with the maximum weight obtained at initialization. Also the cross-section Monte Carlo integration is based on the summed weight. \item If the point is retained, one of the allowed flavour combinations is picked according to the relative weights stored in the full table. \end{Enumerate} Generally, the flavours of the final state are either completely specified by those of the initial state, e.g. as in $\mathrm{q}\mathrm{g} \to \mathrm{q}\mathrm{g}$, or completely decoupled from them, e.g. as in $\mathrm{f}\overline{\mathrm{f}} \to \mathrm{Z}^0 \to \mathrm{f}'\overline{\mathrm{f}}'$. In neither case need therefore the final-state flavours be specified in the cross-section calculation. It is only necessary, in the latter case, to include an overall weight factor, which takes into account the summed contribution of all final states that are to be simulated. For instance, if only the process $\mathrm{Z}^0 \to \e^+\e^-$ is studied, the relevant weight factor is simply $\Gamma_{\mathrm{e}\e} / \Gamma_{\mrm{tot}}$. Once the kinematics and the incoming flavours have been selected, the outgoing flavours can be picked according to the appropriate relative probabilities. In some processes, such as $\mathrm{g}\g \to \mathrm{g}\g$, several different colour flows are allowed, each with its own kinematical dependence of the matrix-element weight, see section \ref{sss:QCDjetclass}. Each colour flow is then given as a separate entry in the table mentioned above, i.e. in total an entry is characterized by the two incoming flavours, a colour-flow index, and the weight. For an accepted phase-space point, the colour flow is selected in the same way as the incoming flavours. The program can also allow the mixed generation of two or more completely different processes, such as $\mathrm{f}\overline{\mathrm{f}} \to \mathrm{Z}^0$ and $\mathrm{q}\overline{\mathrm{q}} \to \mathrm{g}\g$. In that case, each process is initialized separately, with its own set of coefficients $c_i$ and so on. The maxima obtained for the individual cross sections are all expressed in the same units, even when the dimensionality of the phase space is different. (This is because we always transform to a phase space of unit volume, $\int h_{\tau}(\tau) \, \d \tau \equiv 1$, etc.) The above generation scheme need therefore only be generalized as follows: \begin{Enumerate} \item One process is selected among the allowed ones, with a relative probability given by the maximum weight for this process. \item A phase-space point is found, using the distributions $h_{\tau}(\tau)$ and so on, optimized for this particular process. \item The total weight for the phase-space point is evaluated, again with `Jacobians', matrix elements and allowed incoming flavour combinations that are specific to the process. \item The point is retained with a probability given by the ratio of the actual to the maximum weight of the process. If the point is rejected, one has to go back to step 1 and pick a new process. \item Once a phase-space point has been accepted, flavours may be selected, and the event generated in full. \end{Enumerate} It is clear why this works: although phase-space points are selected among the allowed processes according to relative probabilities given by the maximum weights, the probability that a point is accepted is proportional to the ratio of actual to maximum weight. In total, the probability for a given process to be retained is therefore only proportional to the average of the actual weights, and any dependence on the maximum weight is gone. In $\gamma\mathrm{p}$ and $\gamma\gamma$ physics, the different components of the photon give different final states, see section \ref{sss:photoprod}. Technically, this introduces a further level of administration, since each event class contains a set of (partly overlapping) processes. From an ideological point of view, however, it just represents one more choice to be made, that of event class, before the selection of process in step 1 above. When a weighting fails, both class and process have to be picked anew. \subsection{$2 \to 3$ and $2 \to 4$ Processes} The {\tsc{Pythia}} machinery to handle $2 \to 1$ and $2 \to 2$ processes is fairly sophisticated and generic. The same cannot be said about the generation of hard scattering processes with more than two final-state particles. The number of phase-space variables is larger, and it is therefore more difficult to find and transform away all possible peaks in the cross section by a suitably biased choice of phase-space points. In addition, matrix-element expressions for $2 \to 3$ processes are typically fairly lengthy. Therefore {\tsc{Pythia}} only contains a very limited number of $2 \to 3$ and $2 \to 4$ processes, and almost each process is a special case of its own. It is therefore less interesting to discuss details, and we only give a very generic overview. If the Higgs mass is not light, interactions among longitudinal $\mathrm{W}$ and $\mathrm{Z}$ gauge bosons are of interest. In the program, $2 \to 1$ processes such as $\mathrm{W}_{\mrm{L}}^+ \mathrm{W}_{\mrm{L}}^- \to \H^0$ and $2 \to 2$ ones such as $\mathrm{W}_{\mrm{L}}^+ \mathrm{W}_{\mrm{L}}^- \to \mathrm{Z}_{\mrm{L}}^0 \mathrm{Z}_{\mrm{L}}^0$ are included. The former are for use when the $\H^0$ still is reasonably narrow, such that a resonance description is applicable, while the latter are intended for high energies, where different contributions have to be added up. Since the program does not contain $\mathrm{W}_{\mrm{L}}$ or $\mathrm{Z}_{\mrm{L}}$ distributions inside hadrons, the basic hard scattering has to be convoluted with the $\mathrm{q} \to \mathrm{q}' \mathrm{W}_{\mrm{L}}$ and $\mathrm{q} \to \mathrm{q} \mathrm{Z}_{\mrm{L}}$ branchings, to yield effective $2 \to 3$ and $2 \to 4$ processes. However, it is possible to integrate out the scattering angles of the quarks analytically, as well as one energy-sharing variable \cite{Cha85}. Only after an event has been accepted are these other kinematical variables selected. This involves further choices of random variables, according to a separate selection loop. In total, it is therefore only necessary to introduce one additional variable in the basic phase-space selection, which is chosen to be $\hat{s}'$, the squared invariant mass of the full $2 \to 3$ or $2 \to 4$ process, while $\hat{s}$ is used for the squared invariant mass of the inner $2 \to 1$ or $2 \to 2$ process. The $y$ variable is coupled to the full process, since parton-distribution weights have to be given for the original quarks at $x_{1,2} = \sqrt{\tau'} \exp{(\pm y)}$. The $\hat{t}$ variable is related to the inner process, and thus not needed for the $2 \to 3$ processes. The selection of the $\tau' = \hat{s}'/s$ variable is done after $\tau$, but before $y$ has been chosen. To improve the efficiency, the selection is made according to a weighted phase space of the form $\int h_{\tau'}(\tau') \, \d \tau'$, where \begin{equation} h_{\tau'}(\tau') = \frac{c_1}{{\cal I}_1} \frac{1}{\tau'} + \frac{c_2}{{\cal I}_2} \, \frac{(1- \tau / \tau')^3}{\tau'^2} + \frac{c_3}{{\cal I}_3} \, \frac{1}{\tau' (1 - \tau')} ~, \end{equation} in conventional notation. The $c_i$ coefficients are optimized at initialization. The $c_3$ term, peaked at $\tau' \approx 1$, is only used for $\e^+\e^-$ collisions. The choice of $h_{\tau'}$ is roughly matched to the longitudinal gauge-boson flux factor, which is of the form \begin{equation} \left( 1 + \frac{\tau}{\tau'} \right) \, \ln \left( \frac{\tau}{\tau'} \right) - 2 \left( 1 - \frac{\tau}{\tau'} \right) ~. \end{equation} For a light $\H$ the effective $\mathrm{W}$ approximation above breaks down, and it is necessary to include the full structure of the $\mathrm{q} \mathrm{q}' \to \mathrm{q} \mathrm{q}' \H^0$ (i.e. $\mathrm{Z}\Z$ fusion) and $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \mathrm{q}''' \H^0$ (i.e. $\mathrm{W}\W$ fusion) matrix elements. The $\tau'$, $\tau$ and $y$ variables are here retained, and selected according to standard procedures. The Higgs mass is represented by the $\tau$ choice; normally the $\H^0$ is so narrow that the $\tau$ distribution effectively collapses to a $\delta$ function. In addition, the three-body final-state phase space is rewritten as \begin{equation} \left( \prod_{i=3}^5 \frac{1}{(2 \pi)^3} \frac{\d^3 p_i}{2 E_i} \right) \, (2 \pi)^4 \delta^{(4)} (p_3 + p_4 + p_5 - p_1 - p_2) = \frac{1}{(2 \pi)^5} \, \frac{\pi^2}{4 \sqrt{\lambda_{\perp 34}}} \, \d p_{\perp 3}^2 \, \frac{\d \varphi_3}{2 \pi} \, \d p_{\perp 4}^2 \, \frac{\d \varphi_4}{2 \pi} \, \d y_5 ~, \end{equation} where $\lambda_{\perp 34} = (m_{\perp 34}^2 - m_{\perp 3}^2 - m_{\perp 4}^2)^2 - 4 m_{\perp 3}^2 m_{\perp 4}^2$. The outgoing quarks are labelled 3 and 4, and the outgoing Higgs 5. The $\varphi$ angles are selected isotropically, while the two transverse momenta are picked, with some foreknowledge of the shape of the $\mathrm{W} / \mathrm{Z}$ propagators in the cross sections, according to $h_{\perp} (p_{\perp}^2) \, \d p_{\perp}^2$, where \begin{equation} h_{\perp}(p_{\perp}^2) = \frac{c_1}{{\cal I}_1} + \frac{c_2}{{\cal I}_2} \, \frac{1}{m_R^2 + p_{\perp}^2} + \frac{c_3}{{\cal I}_3} \, \frac{1}{(m_R^2 + p_{\perp}^2)^2} ~, \end{equation} with $m_R$ the $\mathrm{W}$ or $\mathrm{Z}$ mass, depending on process, and $c_1 = c_2 = 0.05$, $c_3 = 0.9$. Within the limits given by the other variable choices, the rapidity $y_5$ is chosen uniformly. A final choice remains to be made, which comes from a twofold ambiguity of exchanging the longitudinal momenta of partons 3 and 4 (with minor modifications if they are massive). Here the relative weight can be obtained exactly from the form of the matrix element itself. No good phase-space choice was found for the process $\mathrm{g} \mathrm{g} \to \mathrm{Z}^0 \b \overline{\mathrm{b}}$. This process is therefore not so easy to generate with {\tsc{Pythia}}. What is currently done is to use the basic formalism of $2 \to 2$ processes, where the $\b + \overline{\mathrm{b}}$ system is considered as an effective `resonance'. Two masses are then selected, the $\mathrm{Z}^0$ one according to eq.~(\ref{pg:reshdist}) and the $\b + \overline{\mathrm{b}}$ one according to $\d m^2 / m^2$. Both `decays' are selected isotropically in the respective rest frame, to give the final four fermions in terms of which the matrix element is given. In addition, $\tau$, $y$ and $z$ are selected according to the standard rules for $2 \to 2$ processes. \subsection{Resonance Decays} \label{ss:resdecay} Resonances can be made to decay in two different routines. One is the standard decay treatment (in \ttt{LUDECY}) that can be used for any unstable particle, where decay channels are chosen according to fixed probabilities, and decay angles usually are picked isotropically in the rest frame of the resonance, see section \ref{ss:partdecays}. The more sophisticated treatment (in \ttt{PYRESD}) is the default one for resonances produced in {\tsc{Pythia}}, and is described here. The following are included in the list of resonances: $\mathrm{Z}^0$, $\mathrm{W}^{\pm}$, $\H^0$, $\mathrm{Z}'^0$, $\mathrm{W}'^{\pm}$, $\H'^0$, $\mathrm{A}^0$, $\H^{\pm}$, $\eta_{\mrm{tech}}^0$, $\L_{\mathrm{Q}}$, and $\mathrm{R}^0$. The top is also considered as a resonance if it is assumed to decay before it has time to fragment. Likewise for the fourth generation fermions. If the fourth generation is used to represent excited quarks and leptons, these are also considered to be resonances. \subsubsection{The decay scheme} In the beginning of the decay treatment, either one or two resonances may be present, the former represented by processes such as $\mathrm{q} \overline{\mathrm{q}}' \to \mathrm{W}^+$ and $\mathrm{q} \mathrm{g} \to \mathrm{W}^+ \mathrm{q}'$, the latter by $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{W}^+ \mathrm{W}^-$. If the latter is the case, the decay of the two resonances is considered in parallel (unlike \ttt{LUDECY}, where one particle at a time is made to decay). First the decay channel of each resonance is selected according to the relative weights $H_R^{(f)}$, as described above, evaluated at the actual mass of the resonance, rather than at the nominal one. Threshold factors are therefore fully taken into account, with channels automatically switched off below the threshold. Normally the masses of the decay products are well-defined, but e.g. in decays like $\H^0 \to \mathrm{W}^+ \mathrm{W}^-$ it is also necessary to select the decay product masses. This is done according to two Breit--Wigners of the type in eq.~(\ref{pg:resshapethree}), multiplied by the threshold factor, which depends on both masses. Next the decay angles of the resonance are selected isotropically in its rest frame. Normally the full range of decay angles is available, but in $2 \to 1$ processes the decay angles of the original resonance may be restrained by user cuts, e.g. on the $p_{\perp}$ of the decay products. Based on the angles, the four-momenta of the decay products are constructed and boosted to the correct frame. As a rule, matrix elements are given with quark and lepton masses assumed vanishing. Therefore the four-momentum vectors constructed at this stage are actually massless for all quarks and leptons. The matrix elements may now be evaluated. For a process such as $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{W}^+ \mathrm{W}^- \to \mathrm{e}^+ \nu_{\mathrm{e}} \mu^- \br{\nu}_{\mu}$, the matrix element is a function of the four-momenta of the two incoming fermions and of the four outgoing ones. An upper limit for the event weight can be constructed from the cross section for the basic process $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{W}^+ \mathrm{W}^-$, as already used to select the two $\mathrm{W}$ momenta. If the weighting fails, new resonance decay angles are picked and the procedure is iterated until acceptance. Based on the accepted set of angles, the correct decay product four-momenta are constructed, including previously neglected fermion masses. Quarks and, optionally, leptons are allowed to radiate, using the standard final-state showering machinery, with maximum virtuality given by the resonance mass. In some decays new resonances are produced, and these are then subsequently allowed to decay. Only one resonance pair is considered at a time, i.e. it is not possible to include correlations which involve the simultaneous decay of three or more resonances. This is in fact all that is currently needed: in a process like $\mathrm{q}\overline{\mathrm{q}} \to \mathrm{Z}^0 \H^0 \to \mathrm{Z}^0 \mathrm{W}^+ \mathrm{W}^- \to 6$ fermions, the spinless nature of the $\H^0$ ensures that the $\mathrm{W}^{\pm}$ decays are decoupled from that of the $\mathrm{Z}^0$ (but not from each other). \subsubsection{Cross-section considerations} \label{sss:resdecaycross} The cross section for a process which involves the production of one or several resonances is always reduced to take into account channels not allowed by user flags. This is trivial for a single $s$-channel resonance, cf. eq.~(\ref{pg:Hinoutsym}), but can also be included approximately if several layers of resonance decays are involved. At initialization, the ratio between the user-allowed width and the nominally possible one is evaluated and stored, starting from the lightest resonances and moving upwards. As an example, one first finds the reduction factors for $\mathrm{W}^+$ and for $\mathrm{W}^-$ decays, which need not be the same if e.g. $\mathrm{W}^+$ is allowed to decay only to quarks and $\mathrm{W}^-$ only to leptons. These factors enter together as a weight for the $\H^0 \to \mathrm{W}^+ \mathrm{W}^-$ channel, which is thus reduced in importance compared with other possible Higgs decay channels. This is also reflected in the weight factor of the $\H^0$ itself, where some channels are open in full, others completely closed, and finally some (like the one above) open but with reduced weight. Finally, the weight for the process $\mathrm{q}\overline{\mathrm{q}} \to \mathrm{Z}^0 \H^0$ is evaluated as the product of the $\mathrm{Z}^0$ weight factor and the $\H^0$ one. The standard cross section of the process is multiplied with this weight. Since the restriction on allowed decay modes is already included in the hard process cross section, mixing of different event types is greatly simplified, and the selection of decay channel chains is straightforward. There is a price to be paid, however. The reduction factors evaluated at initialization all refer to resonances at their nominal masses. For instance, the $\mathrm{W}$ reduction factor is evaluated at the nominal $\mathrm{W}$ mass, even when that factor is used, later on, in the description of the decay of a 120 GeV Higgs, where at least one $\mathrm{W}$ would be produced below this mass. We know of no case where this approximation has any serious consequences, however. The weighting procedure works because the number of resonances to be produced, directly or in subsequent decays, can be derived recursively already from the start. It does not work for particles which could also be produced at later stages, such as the parton-shower evolution and the fragmentation. For instance, $\mathrm{D}^0$ mesons can be produced fairly late in the event generation chain, in unknown numbers, and so weights could not be introduced to compensate, e.g. for the forcing of decays only into $\pi^+ \mathrm{K}^-$. For similar reasons the top is only considered as a resonance if it is not allowed to hadronize; see discussion in section \ref{sss:heavflavclass}. One should note that this reduction factor is separate from the description of the resonance shape itself, where the full width of the resonance has to be used. This width is based on the sum of all possible decay modes, not just the simulated ones. {\tsc{Pythia}} does allow the possibility to change also the underlying physics scenario, e.g. to include the decay of a $\mathrm{Z}^0$ into a fourth-generation neutrino. Normally the evaluation of the reduction factors is straightforward. However, for decays into a pair of equal or charge-conjugate resonances, such as $\mathrm{Z}^0 \mathrm{Z}^0$ or $\mathrm{W}^+ \mathrm{W}^-$, it is possible to pick combinations in such a way that the weight of the pair does not factorize into a product of the weight of each resonance itself. To be precise, any decay channel can be given seven different status codes: \begin{Itemize} \item $-1$: a non-existent decay mode, completely switched off and of no concern to us; \item 0: an existing decay channel, which is switched off; \item 1: a channel which is switched on; \item 2: a channel switched on for particles, but off for antiparticles; \item 3: a channel switched on for antiparticles, but off for particles; \item 4: a channel switched on for one of the resonances, but not for both; \item 5: a channel switched on for the other of the resonances, but not for both. \end{Itemize} The meaning of possibilities 4 and 5 is exemplified by the statement `in a $\mathrm{W}^+\mathrm{W}^-$ pair, one $\mathrm{W}$ decays hadronically and the other leptonically', which thus covers the cases where either $\mathrm{W}^+$ or $\mathrm{W}^-$ decays hadronically. Neglecting non-existing channels, each channel belongs to either of the classes above. If we denote the total branching ratio into channels of type $i$ by $r_i$, this then translates into the requirement $r_0 + r_1 + r_2 + r_3 + r_4 + r_5 = 1$. For a single particle the weight factor is $r_1 + r_2 + r_4$, and for a single antiparticle $r_1 + r_3 + r_4$. For a pair of identical resonances, the joint weight is instead \begin{equation} (r_1 + r_2)^2 + 2 (r_1 + r_2) (r_4 + r_5) + 2 r_4 r_5 ~, \end{equation} and for a resonance--antiresonance pair \begin{equation} (r_1 + r_2)(r_1 + r_3) + (2 r_1 + r_2 + r_3) (r_4 + r_5) + 2 r_4 r_5 ~. \label{eq:WWallchancomb} \end{equation} If some channels come with a reduced weight because of restrictions on subsequent decay chains, this may be described in terms of properly reduced $r_i$, so that the sum is less than unity. For instance, in a $\t\overline{\mathrm{t}} \to \b\mathrm{W}^+ \, \overline{\mathrm{b}}\mathrm{W}^-$ process, the $\mathrm{W}$ decay modes may be restricted to $\mathrm{W}^+ \to \mathrm{q}\overline{\mathrm{q}}$ and $\mathrm{W}^- \to \mathrm{e}^-\bar{\nu}_{\mathrm{e}}$, in which case $(\sum r_i)_{\t} \approx 2/3$ and $(\sum r_i)_{\overline{\mathrm{t}}} \approx 1/9$. With index $\pm$ denoting resonance/antiresonance, eq.~(\ref{eq:WWallchancomb}) then generalizes to \begin{equation} (r_1 + r_2)^+ (r_1 + r_3)^- + (r_1 + r_2)^+ (r_4 + r_5)^- + (r_4 + r_5)^+ (r_1 + r_3)^- + r_4^+ r_5^- + r_5^+ r_4^- ~. \label{eq:WWallchancombgen} \end{equation} \subsection{Nonperturbative Processes} \label{ss:nonpertproc} A few processes are not covered by the discussion so far. These are the ones that depend on the details of hadronic wave functions, and therefore are not strictly calculable perturbatively (although perturbation theory may often provide some guidance). What we have primarily in mind is elastic scattering, diffractive scattering and low-$p_{\perp}$ `minimum-bias' events in hadron--hadron collisions, but one can also find corresponding processes in $\gamma \mathrm{p}$ and $\gamma \gamma$ interactions. The description of these processes is rather differently structured from that of the other ones, as is explained below. Models for `minimum-bias' events are discussed in detail in section \ref{ss:multint}, to which we refer for details on this part of the program. \subsubsection{Hadron--hadron interactions} In hadron--hadron interactions, the total hadronic cross section for $AB \to$ anything, $\sigma^{AB}_{\mrm{tot}}$, is calculated using the parametrization of Donnachie and Landshoff \cite{Don92}. In this approach, each cross section appears as the sum of one pomeron term and one reggeon one \begin{equation} \sigma^{AB}_{\mrm{tot}}(s) = X^{AB} \, s^{\epsilon} + Y^{AB} \, s^{-\eta} ~, \label{pg:sigtotpomreg} \end{equation} where $s = E_{\mrm{cm}}^2$. The powers $\epsilon = 0.0808$ and $\eta = 0.4525$ are expected to be universal, whereas the coefficients $X^{AB}$ and $Y^{AB}$ are specific to each initial state. (In fact, the high-energy behaviour given by the pomeron term is expected to be the same for particle and antiparticle interactions, i.e. $X^{\br{A}B} = X^{AB}$.) Parametrizations not provided in \cite{Don92} have been calculated in the same spirit, making use of quark counting rules \cite{Sch93a}. The total cross section is subdivided according to \begin{equation} \sigma^{AB}_{\mrm{tot}}(s) = \sigma^{AB}_{\mrm{el}}(s) + \sigma^{AB}_{\mrm{sd}(XB)}(s) + \sigma^{AB}_{\mrm{sd}(AX)}(s) + \sigma^{AB}_{\mrm{dd}}(s) + \sigma^{AB}_{\mrm{nd}}(s) ~. \label{pg:sigtotsplit} \end{equation} Here `el' is the elastic process $AB \to AB$, `sd$(XB)$' the single diffractive $AB \to XB$, `sd$(AX)$' the single diffractive $AB \to AX$, `dd' the double diffractive $AB \to X_1 X_2$, and `nd' the non-diffractive ones. Higher diffractive topologies, such as central diffraction, are currently neglected. In the following, the elastic and diffractive cross sections and event characteristics are described, as given in the model by Schuler and Sj\"ostrand \cite{Sch94,Sch93a}. The non-diffractive component is identified with the `minimum bias' physics already mentioned, a practical but not unambiguous choice. Its cross section is given by `whatever is left' according to eq.~(\ref{pg:sigtotsplit}), and its properties are discussed in section \ref{ss:multint}. At not too large squared momentum transfers $t$, the elastic cross section can be approximated by a simple exponential fall-off. If one neglects the small real part of the cross section, the optical theorem then gives \begin{equation} \frac{\d\sigma_{\mrm{el}}}{\d t} = \frac{\sigma_{\mrm{tot}}^2}{16 \pi} \, \exp(B_{\mrm{el}} t) ~, \end{equation} and $\sigma_{\mrm{el}} = \sigma_{\mrm{tot}}^2 / 16 \pi B_{\mrm{el}}$. The elastic slope parameter is parametrized by \begin{equation} B_{\mrm{el}} = B^{AB}_{\mrm{el}}(s) = 2 b_A + 2 b_B + 4 s^{\epsilon} -4.2 ~, \end{equation} with $s$ given in units of GeV and $B_{\mrm{el}}$ in GeV$^{-2}$. The constants $b_{A,B}$ are $b_{\mathrm{p}} = 2.3$, $b_{\pi,\rho,\omega,\phi} = 1.4$, $b_{\mathrm{J}/\psi} = 0.23$. The increase of the slope parameter with c.m. energy is faster than the logarithmically one conventionally assumed; that way the ratio $\sigma_{\mrm{el}} / \sigma_{\mrm{tot}}$ remains well-behaved at large energies. The diffractive cross sections are given by \begin{eqnarray} \frac{\d\sigma_{\mrm{sd}(XB)}(s)}{\d t \, \d M^2} & = & \frac{g_{3\mathrm{I}\!\mathrm{P}}}{16\pi} \, \beta_{A\mathrm{I}\!\mathrm{P}} \, \beta_{B\mathrm{I}\!\mathrm{P}}^2 \, \frac{1}{M^2} \, \exp(B_{\mrm{sd}(XB)}t) \, F_{\mrm{sd}} ~, \nonumber \\ \frac{\d\sigma_{\mrm{sd}(AX)}(s)}{\d t \, \d M^2} & = & \frac{g_{3\mathrm{I}\!\mathrm{P}}}{16\pi} \, \beta_{A\mathrm{I}\!\mathrm{P}}^2 \, \beta_{B\mathrm{I}\!\mathrm{P}} \, \frac{1}{M^2} \, \exp(B_{\mrm{sd}(AX)}t) \, F_{\mrm{sd}} ~, \nonumber \\ \frac{\d\sigma_{\mrm{dd}}(s)}{\d t \, \d M_1^2 \, \d M_2^2} & = & \frac{g_{3\mathrm{I}\!\mathrm{P}}^2}{16\pi} \, \beta_{A\mathrm{I}\!\mathrm{P}} \, \beta_{B\mathrm{I}\!\mathrm{P}} \, \frac{1}{M_1^2} \, \frac{1}{M_2^2} \, \exp(B_{\mrm{dd}}t) \, F_{\mrm{dd}} ~. \end{eqnarray} The couplings $\beta_{A\mathrm{I}\!\mathrm{P}}$ are related to the pomeron term $X^{AB} s^{\epsilon}$ of the total cross section parametrization, eq.~(\ref{pg:sigtotpomreg}). Picking a reference scale $\sqrt{s_{\mrm{ref}}} = 20$ GeV, the couplings are given by $\beta_{A\mathrm{I}\!\mathrm{P}}\beta_{B\mathrm{I}\!\mathrm{P}} = X^{AB} \, s_{\mrm{ref}}^{\epsilon}$. The triple-pomeron coupling is determined from single-diffractive data to be $g_{3\mathrm{I}\!\mathrm{P}} \approx 0.318$ mb$^{1/2}$; within the context of the formulae in this section. The spectrum of diffractive masses $M$ is taken to begin 0.28 GeV $\approx 2 m_{\pi}$ above the mass of the respective incoming particle and extend to the kinematical limit. The simple $\d M^2 / M^2$ form is modified by the mass-dependence in the diffractive slopes and in the $F_{\mrm{sd}}$ and $F_{\mrm{dd}}$ factors. The slope parameters are assumed to be \begin{eqnarray} B_{\mrm{sd}(XB)}(s) & = & 2b_B + 2\alpha' \ln\left(\frac{s}{M^2}\right) ~, \nonumber \\ B_{\mrm{sd}(AX)}(s) & = & 2b_A + 2\alpha' \ln\left(\frac{s}{M^2}\right) ~, \nonumber \\ B_{\mrm{dd}}(s) & = & 2\alpha' \ln\left(e^4 + \frac{s s_0}{M_1^2 M_2^2} \right) ~. \end{eqnarray} Here $\alpha' = 0.25$ GeV$^{-2}$ and conventionally $s_0$ is picked as $s_0 = 1 / \alpha'$. The term $e^4$ in $B_{\mrm{dd}}$ is added by hand to avoid a breakdown of the standard expression for large values of $M_1^2 M_2^2$. The $b_{A,B}$ terms protect $B_{\mrm{sd}}$ from breaking down; however a minimum value of 2 is still explicitly required for $B_{\mrm{sd}}$, which comes into play e.g. for a $\mathrm{J}/\psi$ state (as part of a VMD photon beam). The kinematical range in $t$ depends on all the masses of the problem. In terms of the scaled variables $\mu_1 = m_A^2/s$, $\mu_2 = m_B^2/s$, $\mu_3 = M_{(1)}^2/s$ ($=m_A^2/s$ when $A$ scatters elastically), $\mu_4 = M_{(2)}^2/s$ ($=m_B^2/s$ when $B$ scatters elastically), and the combinations \begin{eqnarray} C_1 & = & 1 - (\mu_1 + \mu_2 + \mu_3 + \mu_4) + (\mu_1 - \mu_2) (\mu_3 - \mu_4) ~, \nonumber \\ C_2 & = & \sqrt{(1 - \mu_1 -\mu_2)^2 - 4 \mu_1 \mu_2} \, \sqrt{(1 - \mu_3 - \mu_4)^2 - 4 \mu_3 \mu_4} ~, \nonumber \\ C_3 & = & (\mu_3 - \mu_1) (\mu_4 - \mu_2) + (\mu_1 + \mu_4 - \mu_2 - \mu_3) (\mu_1 \mu_4 - \mu_2 \mu_3) ~, \end{eqnarray} one has $t_{\mathrm{min}} < t < t_{\mathrm{max}}$ with \begin{eqnarray} t_{\mathrm{min}} & = & - \frac{s}{2} (C_1 + C_2) ~, \nonumber \\ t_{\mathrm{max}} & = & - \frac{s}{2} (C_1 - C_2) = - \frac{s}{2} \, \frac{4C_3}{C_1 + C_2} = \frac{s^2 C_3}{t_{\mathrm{min}}} ~. \end{eqnarray} The Regge formulae above for single- and double-diffractive events are supposed to hold in certain asymptotic regions of the total phase space. Of course, there will be diffraction also outside these restrictive regions. Lacking a theory which predicts differential cross sections at arbitrary $t$ and $M^2$ values, the Regge formulae are used everywhere, but fudge factors are introduced in order to obtain `sensible' behaviour in the full phase space. These factors are: \begin{eqnarray} F_{\mrm{sd}} & = & \left( 1 - \frac{M^2}{s} \right) \left( 1 + \frac{c_{\mrm{res}} \, M_{\mrm{res}}^2} {M_{\mrm{res}}^2 + M^2} \right) ~, \nonumber \\ F_{\mrm{dd}} & = & \left( 1 - \frac{\left( M_1 + M_2 \right)^2}{s} \right) \left( \frac{s\, m_{\mathrm{p}}^2}{ s\, m_{\mathrm{p}}^2 + M_1^2\, M_2^2} \right) \nonumber \\ & \times & \left( 1 + \frac{c_{\mrm{res}} \, M_{\mrm{res}}^2} {M_{\mrm{res}}^2 + M_1^2} \right) \left( 1 + \frac{c_{\mrm{res}} \, M_{\mrm{res}}^2} {M_{\mrm{res}}^2 + M_2^2} \right) ~. \end{eqnarray} The first factor in either expression suppresses production close to the kinematical limit. The second factor in $F_{dd}$ suppresses configurations where the two diffractive systems overlap in rapidity space. The final factors give an enhancement of the low-mass region, where a resonance structure is observed in the data. Clearly a more detailed modelling would have to be based on a set of exclusive states rather than on this smeared-out averaging procedure. A reasonable fit to $\mathrm{p}\p / \overline{\mathrm{p}}\mathrm{p}$ data is obtained for $c_{\mrm{res}} = 2$ and $M_{\mrm{res}} = 2$~GeV, for an arbitary particle $A$ which is diffractively excited we use $M_{\mrm{res}}^A = m_A - m_{\mathrm{p}} + 2$~GeV. The diffractive cross-section formulae above have been integrated for a set of c.m. energies, starting at 10 GeV, and the results have been parametrized. The form of these parametrizations is given in ref. \cite{Sch94}, with explicit numbers for the $\mathrm{p}\p/\overline{\mathrm{p}}\mathrm{p}$ case. {\tsc{Pythia}} also contains similar parametrizations for $\pi\mathrm{p}$ (assumed to be same as $\rho\mathrm{p}$ and $\omega\mathrm{p}$), $\phi\mathrm{p}$, $\mathrm{J}/\psi\mathrm{p}$, $\rho\rho$ ($\pi\pi$ etc.), $\rho\phi$, $\rho\mathrm{J}/\psi$, $\phi\phi$, $\phi\mathrm{J}/\psi$ and $\mathrm{J}/\psi\Jpsi$. The processes above do not obey the ordinary event mixing strategy. First of all, since their total cross sections are known, it is possible to pick the appropriate process from the start, and then remain with that choice. In other words, if the selection of kinematical variables fails, one would not go back and pick a new process, the way it was done in section \ref{sss:mixingproc}. Second, it is not possible to impose any cuts or restrain allowed incoming or outgoing flavours: if not additional information were to be provided, it would make the whole scenario ill-defined. Third, it is not recommended to mix generation of these processes with that of any of the other ones: normally the other processes have so small cross sections that they would almost never be generated anyway. (We here exclude the cases of `underlying events' and `pile-up events', where mixing is provided for, and even is a central part of the formalism, see sections \ref{ss:multint} and \ref{ss:pileup}.) Once the cross-section parametrizations has been used to pick one of the processes, the variables $t$ and $M$ are selected according to the formulae given above. A $\rho^0$ formed by $\gamma \to \rho^0$ in elastic or diffractive scattering is polarized, and therefore its decay angular distribution in $\rho^0 \to \pi^+ \pi^-$ is taken to be proportional to $\sin^2 \theta$, where the reference axis is given by the $\rho^0$ direction of motion. A light diffractive system, with a mass less than 1 GeV above the mass of the incoming particle, is allowed to decay isotropically into a two-body state. Single-resonance diffractive states, such as a $\Delta^+$, are therefore not explicily generated, but are assumed described in an average, smeared-out sense. A more massive diffractive system is subsequently treated as a string with the quantum numbers of the original hadron. Since the exact nature of the pomeron exchanged between the hadrons is unknown, two alternatives are included. In the first, the pomeron is assumed to couple to (valence) quarks, so that the string is stretched directly between the struck quark and the remnant diquark (antiquark) of the diffractive state. In the second, the interaction is rather with a gluon, giving rise to a `hairpin' configuration in which the string is stretched from a quark to a gluon and then back to a diquark (antiquark). Both of these scenarios could be present in the data; the default choice is to mix them in equal proportions. There is experimental support for more complicated scenarios \cite{Ing85}, wherein the pomeron has a partonic substructure, which e.g. can lead to high-$p_{\perp}$ jet production in the diffractive system. The full machinery, wherein a pomeron spectrum is folded with a pomeron-proton hard interaction, is not available in {\tsc{Pythia}}. \subsubsection{Photoproduction and $\gamma\gamma$ physics} \label{sss:photoprod} The photoproduction part is still under active development. Currently only interactions between a hadron and a real photon have been studied in detail. $\gamma\gamma$ physics is under study \cite{Sch94a}, and is now preliminarily included for real photons. Deep inelastic scattering on a real photon is also preliminarily included. In the future it is hoped to add interactions of mildly virtual photons (the transition region between real photons and deep inelastic scattering). The total $\gamma\mathrm{p}$ and $\gamma\gamma$ cross sections can again be parametrized in a form like eq.~(\ref{pg:sigtotpomreg}), which is not so obvious since the photon has more complicated structure than an ordinary hadron. In fact, the structure is still not so well understood. The model we outline is the one studied by Schuler and Sj\"ostrand \cite{Sch93,Sch93a}. In this model the physical photon is represented by \begin{equation} | \gamma \rangle = \sqrt{Z_3} \, | \gamma_B \rangle + \sum_{V=\rho^0,\omega,\phi,\mathrm{J}/\psi} \frac{e}{f_V} \, | V \rangle + \frac{e}{f_{\mathrm{q}\overline{\mathrm{q}}}} \, | \mathrm{q}\overline{\mathrm{q}} \rangle + \sum_{\ell=\mathrm{e},\mu,\tau} \frac{e}{f_{\ell\ell}} \, | \ell^+ \ell^- \rangle ~. \label{pg:gammadecompo} \end{equation} By virtue of this superposition, one is led to a model of $\gamma\mathrm{p}$ interactions, where three different kinds of events may be distinguished: \begin{Itemize} \item Direct events, wherein the bare photon $| \gamma_B \rangle$ interacts directly with a parton from the proton. The process is perturbatively calculable, and no parton distributions of the photon are involved. The typical event structure is two high-$p_{\perp}$ jets and a proton remnant, while the photon does not leave behind any remnant. \item VMD events, in which the photon fluctuates into a vector meson, predominantly a $\rho^0$. All the event classes known from ordinary hadron--hadron interactions may thus occur here, such as elastic, diffractive, low-$p_{\perp}$ and high-$p_{\perp}$ events. For the latter, one may define (VMD) parton distributions of the photon, and the photon also leaves behind a beam remnant. This remnant is smeared in transverse momentum by a typical `primordial $k_{\perp}$' of a few hundred MeV. \item Anomalous events, in which the photon fluctuates into a $\mathrm{q}\overline{\mathrm{q}}$ pair of larger virtuality than in the VMD class. This process is perturbatively calculable, as is the subsequent QCD evolution. It gives rise to the so-called anomalous part of the parton distributions ofthe photon, whence the name for the class. It is assumed that only high-$p_{\perp}$ events may occur. Either the $\mathrm{q}$ or the $\overline{\mathrm{q}}$ plays the r\^ole of a beam remnant, but this remnant has a larger $p_{\perp}$ than in the VMD case, related to the virtuality of the $\gamma \leftrightarrow \mathrm{q}\overline{\mathrm{q}}$ fluctuation. \end{Itemize} The $| \ell^+ \ell^- \rangle$ states can only interact strongly with partons inside the hadron at higher orders, and can therefore be neglected. In order that the above classification is smooth and free of double counting, one has to introduce scales that separate the three components. The main one is $p_0$, which separates the low-mass vector meson region from the high-mass $| \mathrm{q}\overline{\mathrm{q}} \rangle$ one, $p_0 \approx m_{\phi}/2 \approx 0.5$ GeV. Since it is the same $\gamma\mathrm{q}\overline{\mathrm{q}}$ vertex that is responsible for the bare $\gamma \mathrm{p}$ interactions, $p_0$ is also the lower cut-off of the photon--parton cross sections. In addition, a $p_{\perp\mathrm{min}}$ cut-off is needed to separate low-$p_{\perp}$ and high-$p_{\perp}$ physics; see section \ref{ss:multint}. As it turns out, somewhat different $p_{\perp\mathrm{min}}$ values are needed for the VMD and anomalous parts; at least qualitatively this can be understood in terms of different sizes of the wave functions. The VMD and anomalous events are together called resolved ones. In terms of high-$p_{\perp}$ jet production, the VMD and anomalous contributions can be combined into a total resolved one, and the same for parton-distribution functions. However, the two classes differ in the structure of the underlying event and in the appearance of soft processes. In terms of cross sections, eq.\ (\ref{pg:gammadecompo}) corresponds to \begin{equation} \sigma_{\mrm{tot}}^{\gamma\mathrm{p}}(s) = \sigma_{\mrm{dir}}^{\gamma\mathrm{p}}(s) + \sigma_{\mrm{VMD}}^{\gamma\mathrm{p}}(s) + \sigma_{\mrm{anom}}^{\gamma\mathrm{p}}(s) ~. \label{pg:gammacrossdecompo} \end{equation} The direct cross section is, to lowest order, the perturbative cross section for the two processes $\gamma\mathrm{q} \to \mathrm{q}\mathrm{g}$ and $\gamma\mathrm{g} \to \mathrm{q}\overline{\mathrm{q}}$, with a lower cut-off $p_{\perp} > p_0$. Properly speaking, this should be multiplied by the $Z_3$ coefficient, \begin{equation} Z_3 = 1 - \sum_{V=\rho^0,\omega,\phi,\mathrm{J}/\psi} \left( \frac{e}{f_V} \right)^2 - \left( \frac{e}{f_{\mathrm{q}\overline{\mathrm{q}}}} \right)^2 - \sum_{\ell=\mathrm{e},\mu,\tau} \left( \frac{e}{f_{\ell\ell}} \right)^2 ~, \end{equation} but normally $Z_3$ is so close to unity as to make no difference. The VMD factor $(e/f_V)^2 = 4\pi\alpha_{\mathrm{em}}/f_V^2$ gives the probability for the transition $\gamma \to V$. The coefficients $f_V^2/4\pi$ are determined from data to be (with a non-negligible amount of uncertainty) 2.20 for $\rho^0$, 23.6 for $\omega$, 18.4 for $\phi$ and 11.5 for $\mathrm{J}/\psi$. Together these numbers imply that the photon can be found in a VMD state about 0.4\% of the time, dominated by the $\rho^0$ contribution. All the properties of the VMD interactions can be obtained by appropriately scaling down $V\mathrm{p}$ physics predictions. Thus the whole machinery developed in the previous subsection for hadron--hadron interactions is directly applicable. Also parton distributions of the VMD component inside the photon are obtained by suitable rescaling. The contribution from the `anomalous' high-mass fluctuations depends on the typical scale $\mu$ of the interaction \begin{equation} \left( \frac{e}{f_{\mathrm{q}\overline{\mathrm{q}}}} \right)^2 \approx \frac{\alpha_{\mathrm{em}}}{2\pi} \, \frac{N_C}{3} \, \left( 2 \sum_{\mathrm{q}} e_{\mathrm{q}}^2 \right) \ln\left( \frac{\mu^2}{p_0^2} \right) ~, \label{anomintfirst} \end{equation} where $N_C = 3$ and $\mathrm{q}$ runs over the quarks that can be taken massless compared with $\mu$. The logarithmic increase with $\mu$ implies that the anomalous contribution to the total photoproduction cross section ($\mu \sim m_V$) is less important than that to high-$p_{\perp}$ jet production ($\mu \sim p_{\perp}$). To first approximation, therefore only perturbative jet production above some $p_{\perp\mathrm{min}}$ scale is considered. This includes the standard QCD parton--parton scattering processes, with anomalous-photon parton distributions that are fully perturbatively calculable \cite{Sch95}. In order to satisfy the equality in eq.~(\ref{pg:gammacrossdecompo}), with the total cross section known and the direct and VDM contributions already fixed, a behaviour roughly like \begin{equation} p_{\perp\mathrm{min}}^{\mrm{anom}}(s) = 0.70 + 0.17 \log^2(1.+0.05\sqrt{s}) \end{equation} is needed over the HERA energy range. This is to be seen entirely as a pragmatic parametrization, not be given any fundamental interpretation. It is based on SaS set~1D, another set might well require a somewhat different form. In $\gamma\gamma$ physics \cite{Sch94a}, the superposition in eq.~(\ref{pg:gammadecompo}) applies separately for each of the two incoming photons. In total there are therefore $3 \times 3 = 9$ combinations. However, trivial symmetry reduces this to six distinct classes, written in terms of the total cross section (cf. eq.~(\ref{pg:gammacrossdecompo})) as \begin{eqnarray} \sigma_{\mrm{tot}}^{\gamma\gamma}(s) & = & \sigma_{\mrm{dir}\times\mrm{dir}}^{\gamma\gamma}(s) + \sigma_{\mrm{VMD}\times\mrm{VMD}}^{\gamma\gamma}(s) + \sigma_{\mrm{anom}\times\mrm{anom}}^{\gamma\gamma}(s) \nonumber \\ & + & 2 \sigma_{\mrm{dir}\times\mrm{VMD}}^{\gamma\gamma}(s) + 2 \sigma_{\mrm{dir}\times\mrm{anom}}^{\gamma\gamma}(s) + 2 \sigma_{\mrm{VMD}\times\mrm{anom}}^{\gamma\gamma}(s) ~. \label{pg:gagacrossdecompo} \end{eqnarray} A parametrization of the total $\gamma\gamma$ cross section and comments on its subdivision into the six classes is found in \cite{Sch94a}. The six different kinds of $\gamma\gamma$ events are thus: \begin{Itemize} \item The direct$\times$direct events, which correspond to the subprocess $\gamma\gamma \to \mathrm{q}\overline{\mathrm{q}}$ (or $\ell^+\ell^-$). The typical event structure is two high-$p_{\perp}$ jets and no beam remnants. The lower cut-off is $p_{\perp} > p_0$. \item The VMD$\times$VMD events, which have the same properties as the VMD $\gamma\mathrm{p}$ events. There are four by four combinations of the two incoming vector mesons, with one VMD factor for each meson. \item The anomalous$\times$anomalous events, wherein each photon fluctuates into a $\mathrm{q}\overline{\mathrm{q}}$ pair of larger virtuality than in the VMD class. One parton of each pair gives a beam remnant, whereas the other (or a daughter parton thereof) participates in a high-$p_{\perp}$ scattering, with $p_{\perp} > p_{\perp\mathrm{min}}^{\mrm{anom}}$. \item The direct$\times$VMD events, which have the same properties as the direct $\gamma\mathrm{p}$ events. \item The direct$\times$anomalous events, in which a bare photon interacts with a parton from the anomalous photon. The lower cut-off for the hard scattering is given by $p_{\perp\mathrm{min}}^{\mrm{anom}}$. The typical structure is then two high-$p_{\perp}$ jets and a beam remnant. \item The VMD$\times$anomalous events, which have the same properties as the anomalous $\gamma\mathrm{p}$ events. \end{Itemize} In much of the literature, where a coarser classification us used, our direct$\times$direct is called direct, our direct$\times$VMD and direct$\times$anomalous is called 1-resolved since they both involve one resolved photon which gives a beam remnant, and the rest are called 2-resolves since both photons are resolved and give beam remnants. \clearpage \section{Physics Processes in PYTHIA} \label{s:pytproc} In this section we enumerate the physics processes that are available in {\tsc{Pythia}}, introducing the ISUB code that can be used to select desired processes. A number of comments are made about the physics scenarios involved, in particular with respect to underlying assumptions and domain of validity. The section closes with a survey of interesting processes by machine. \subsection{The Process Classification Scheme} \label{ss:ISUBcode} A wide selection of fundamental $2 \to 1$ and $2 \to 2$ tree processes of the Standard Model (electroweak and strong) has been included in {\tsc{Pythia}}, and slots are provided for many more, not yet implemented. In addition, a few `minimum-bias'-type processes (like elastic scattering), loop graphs, box graphs, $2 \to 3$ tree graphs and some non-Standard Model processes are included. The classification is not always unique. A process that proceeds only via an $s$-channel state is classified as a $2 \to 1$ process (e.g. $\mathrm{q} \overline{\mathrm{q}} \to \gamma^* / \Z^0 \to \e^+\e^-$), but a $2 \to 2$ cross section may well have contributions from $s$-channel diagrams ($\mathrm{g} \mathrm{g} \to \mathrm{g} \mathrm{g}$ obtains contributions from $\mathrm{g} \mathrm{g} \to \mathrm{g}^* \to \mathrm{g} \mathrm{g}$). Also, in the program, $2 \to 1$ and $2 \to 2$ graphs may sometimes be folded with two $1 \to 2$ splittings to form effective $2 \to 3$ or $2 \to 4$ processes ($\mathrm{W}^+ \mathrm{W}^- \to \H^0$ is folded with $\mathrm{q} \to \mathrm{q}'' \mathrm{W}^+$ and $\mathrm{q}' \to \mathrm{q}''' \mathrm{W}^-$ to give $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \mathrm{q}''' \H^0$). It is possible to select a combination of subprocesses to simulate, and also afterwards to know which subprocess was actually selected in each event. For this purpose, all subprocesses are numbered according to an ISUB code. The list of possible codes is given in Tables \ref{t:procone}, \ref{t:proctwo}, \ref{t:procthree} and \ref{t:procfour}. Only processes marked with a `+' sign in the first column have been implemented in the program to date. Although ISUB codes were originally designed in a logical fashion, we must admit that subsequent developments of the program have tended to obscure the structure. For instance, the process numbers for Higgs production are spread out, in part as a consequence of the original classification, in part because further production mechanisms have been added one at a time, in whatever free slots could be found. At some future date the subprocess list will therefore be reorganized. In the thematic descriptions that follow the main tables, the processes of interest are repeated in a more logical order. If you want to look for a specific process, it will be easier to find it there. \begin{table}[pt] \caption{Subprocess codes, part 1. First column is `+' for processes implemented and blank for those that are only foreseen. Second is the subprocess number \ttt{ISUB}, and third the description of the process. The final column gives references from which the cross sections have been obtained. See text for further information. \protect\label{t:procone} } \begin{center} \begin{tabular}{|c|r|l|l|@{\protect\rule{0mm}{\tablinsep}}} \hline In & No. & Subprocess & Reference \\ \hline & & a) $2 \to 1$, tree & \\ + & 1 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma^* / \Z^0$ & \cite{Eic84} \\ + & 2 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}^+$ & \cite{Eic84} \\ + & 3 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \H^0$ & \cite{Eic84} \\ & 4 & $\gamma \mathrm{W}^+ \to \mathrm{W}^+$ & \\ + & 5 & $\mathrm{Z}^0 \mathrm{Z}^0 \to \H^0$ & \cite{Eic84,Cha85} \\ & 6 & $\mathrm{Z}^0 \mathrm{W}^+ \to \mathrm{W}^+$ & \\ & 7 & $\mathrm{W}^+ \mathrm{W}^- \to \mathrm{Z}^0$ & \\ + & 8 & $\mathrm{W}^+ \mathrm{W}^- \to \H^0$ & \cite{Eic84,Cha85} \\ \hline & & b) $2 \to 2$, tree & \\ + & 10 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_i \mathrm{f}_j$ (QFD) & \cite{Ing87b} \\ + & 11 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_i \mathrm{f}_j$ (QCD) & \cite{Com77,Ben84,Eic84,Chi90} \\ + & 12 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \cite{Com77,Ben84,Eic84,Chi90} \\ + & 13 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{g} \mathrm{g}$ & \cite{Com77,Ben84} \\ + & 14 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{g} \gamma$ & \cite{Hal78,Ben84} \\ + & 15 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{g} \mathrm{Z}^0$ & \cite{Eic84} \\ + & 16 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{g} \mathrm{W}^+$ & \cite{Eic84} \\ & 17 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{g} \H^0$ & \\ + & 18 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma \gamma$ & \cite{Ber84} \\ + & 19 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma \mathrm{Z}^0$ & \cite{Eic84} \\ + & 20 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \gamma \mathrm{W}^+$ & \cite{Eic84,Sam91} \\ & 21 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma \H^0$ & \\ + & 22 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{Z}^0 \mathrm{Z}^0$ & \cite{Eic84,Gun86} \\ + & 23 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{Z}^0 \mathrm{W}^+$ & \cite{Eic84,Gun86} \\ + & 24 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{Z}^0 \H^0$ & \cite{Ber85} \\ + & 25 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{W}^+ \mathrm{W}^-$ & \cite{Bar94,Gun86} \\ + & 26 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}^+ \H^0$ & \cite{Eic84} \\ & 27 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \H^0 \H^0$ & \\ + & 28 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_i \mathrm{g}$ & \cite{Com77,Ben84} \\ + & 29 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_i \gamma$ & \cite{Hal78,Ben84} \\ + & 30 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_i \mathrm{Z}^0$ & \cite{Eic84} \\ + & 31 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_k \mathrm{W}^+$ & \cite{Eic84} \\ & 32 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_i \H^0$ & \\ + & 33 & $\mathrm{f}_i \gamma \to \mathrm{f}_i \mathrm{g}$ & \cite{Duk82} \\ + & 34 & $\mathrm{f}_i \gamma \to \mathrm{f}_i \gamma$ & \cite{Duk82} \\ + & 35 & $\mathrm{f}_i \gamma \to \mathrm{f}_i \mathrm{Z}^0$ & \cite{Gab86} \\ + & 36 & $\mathrm{f}_i \gamma \to \mathrm{f}_k \mathrm{W}^+$ & \cite{Gab86} \\ & 37 & $\mathrm{f}_i \gamma \to \mathrm{f}_i \H^0$ & \\ & 38 & $\mathrm{f}_i \mathrm{Z}^0 \to \mathrm{f}_i \mathrm{g}$ & \\ & 39 & $\mathrm{f}_i \mathrm{Z}^0 \to \mathrm{f}_i \gamma$ & \\ & 40 & $\mathrm{f}_i \mathrm{Z}^0 \to \mathrm{f}_i \mathrm{Z}^0$ & \\ \hline \end{tabular} \end{center} \end{table} \begin{table}[pt] \caption{Subprocess codes, part 2. First column is `+' for processes implemented and blank for those that are only foreseen. Second is the subprocess number \ttt{ISUB}, and third the description of the process. The final column gives references from which the cross sections have been obtained. See text for further information. \protect\label{t:proctwo} } \begin{center} \begin{tabular}{|c|r|l|l|@{\protect\rule{0mm}{\tablinsep}}} \hline In & No. & Subprocess & Reference \\ \hline & & b) $2 \to 2$, tree (cont'd) & \\ & 41 & $\mathrm{f}_i \mathrm{Z}^0 \to \mathrm{f}_k \mathrm{W}^+$ & \\ & 42 & $\mathrm{f}_i \mathrm{Z}^0 \to \mathrm{f}_i \H^0$ & \\ & 43 & $\mathrm{f}_i \mathrm{W}^+ \to \mathrm{f}_k \mathrm{g}$ & \\ & 44 & $\mathrm{f}_i \mathrm{W}^+ \to \mathrm{f}_k \gamma$ & \\ & 45 & $\mathrm{f}_i \mathrm{W}^+ \to \mathrm{f}_k \mathrm{Z}^0$ & \\ & 46 & $\mathrm{f}_i \mathrm{W}^+ \to \mathrm{f}_k \mathrm{W}^+$ & \\ & 47 & $\mathrm{f}_i \mathrm{W}^+ \to \mathrm{f}_k \H^0$ & \\ & 48 & $\mathrm{f}_i \H^0 \to \mathrm{f}_i \mathrm{g}$ & \\ & 49 & $\mathrm{f}_i \H^0 \to \mathrm{f}_i \gamma$ & \\ & 50 & $\mathrm{f}_i \H^0 \to \mathrm{f}_i \mathrm{Z}^0$ & \\ & 51 & $\mathrm{f}_i \H^0 \to \mathrm{f}_k \mathrm{W}^+$ & \\ & 52 & $\mathrm{f}_i \H^0 \to \mathrm{f}_i \H^0$ & \\ + & 53 & $\mathrm{g} \mathrm{g} \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \cite{Com77,Ben84} \\ + & 54 & $\mathrm{g} \gamma \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \cite{Duk82} \\ & 55 & $\mathrm{g} \mathrm{Z}^0 \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \\ & 56 & $\mathrm{g} \mathrm{W}^+ \to \mathrm{f}_k \overline{\mathrm{f}}_l$ & \\ & 57 & $\mathrm{g} \H^0 \to \mathrm{f}_k \overline{\mathrm{f}}_l$ & \\ + & 58 & $\gamma \gamma \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \cite{Bar90} \\ & 59 & $\gamma \mathrm{Z}^0 \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \\ & 60 & $\gamma \mathrm{W}^+ \to \mathrm{f}_k \overline{\mathrm{f}}_l$ & \\ & 61 & $\gamma \H^0 \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \\ & 62 & $\mathrm{Z}^0 \mathrm{Z}^0 \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \\ & 63 & $\mathrm{Z}^0 \mathrm{W}^+ \to \mathrm{f}_k \overline{\mathrm{f}}_l$ & \\ & 64 & $\mathrm{Z}^0 \H^0 \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \\ & 65 & $\mathrm{W}^+ \mathrm{W}^- \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \\ & 66 & $\mathrm{W}^+ \H^0 \to \mathrm{f}_k \overline{\mathrm{f}}_l$ & \\ & 67 & $\H^0 \H^0 \to \mathrm{f}_k \overline{\mathrm{f}}_k$ & \\ + & 68 & $\mathrm{g} \mathrm{g} \to \mathrm{g} \mathrm{g}$ & \cite{Com77,Ben84} \\ + & 69 & $\gamma \gamma \to \mathrm{W}^+ \mathrm{W}^-$ & \cite{Kat83} \\ + & 70 & $\gamma \mathrm{W}^+ \to \mathrm{Z}^0 \mathrm{W}^+$ & \cite{Kun87} \\ + & 71 & $\mathrm{Z}^0 \mathrm{Z}^0 \to \mathrm{Z}^0 \mathrm{Z}^0$ (longitudinal) & \cite{Abb87} \\ + & 72 & $\mathrm{Z}^0 \mathrm{Z}^0 \to \mathrm{W}^+ \mathrm{W}^-$ (longitudinal) & \cite{Abb87} \\ + & 73 & $\mathrm{Z}^0 \mathrm{W}^+ \to \mathrm{Z}^0 \mathrm{W}^+$ (longitudinal) & \cite{Dob91} \\ & 74 & $\mathrm{Z}^0 \H^0 \to \mathrm{Z}^0 \H^0$ & \\ & 75 & $\mathrm{W}^+ \mathrm{W}^- \to \gamma \gamma$ & \\ + & 76 & $\mathrm{W}^+ \mathrm{W}^- \to \mathrm{Z}^0 \mathrm{Z}^0$ (longitudinal) & \cite{Ben87b} \\ + & 77 & $\mathrm{W}^+ \mathrm{W}^{\pm} \to \mathrm{W}^+ \mathrm{W}^{\pm}$ (longitudinal) & \cite{Dun86,Bar90a} \\ & 78 & $\mathrm{W}^+ \H^0 \to \mathrm{W}^+ \H^0$ & \\ & 79 & $\H^0 \H^0 \to \H^0 \H^0$ & \\ + & 80 & $\mathrm{q}_i \gamma \to \mathrm{q}_k \pi^{\pm}$ & \cite{Bag82} \\ \hline \end{tabular} \end{center} \end{table} \begin{table}[pt] \caption{Subprocess codes, part 3. First column is `+' for processes implemented and blank for those that are only foreseen. Second is the subprocess number \ttt{ISUB}, and third the description of the process. The final column gives references from which the cross sections have been obtained. See text for further information. \protect\label{t:procthree} } \begin{center} \begin{tabular}{|c|r|l|l|@{\protect\rule{0mm}{\tablinsep}}} \hline In & No. & Subprocess & Reference \\ \hline & & c) $2 \to 2$, tree, massive final quarks & \\ + & 81 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{Q}_k \overline{\mathrm{Q}}_k$ & \cite{Com79} \\ + & 82 & $\mathrm{g} \mathrm{g} \to \mathrm{Q}_k \overline{\mathrm{Q}}_k$ & \cite{Com79} \\ + & 83 & $\mathrm{q}_i \mathrm{f}_j \to \mathrm{Q}_k \mathrm{f}_l$ & \cite{Dic86} \\ + & 84 & $\mathrm{g} \gamma \to \mathrm{Q}_k \overline{\mathrm{Q}}_k$ & \cite{Fon81} \\ + & 85 & $\gamma \gamma \to \mathrm{F}_k \overline{\mathrm{F}}_k$ & \cite{Bar90} \\ + & 86 & $\mathrm{g} \mathrm{g} \to \mathrm{J}/\psi \mathrm{g}$ & \cite{Bai83} \\ + & 87 & $\mathrm{g} \mathrm{g} \to \chi_{0 \c} \mathrm{g}$ & \cite{Gas87} \\ + & 88 & $\mathrm{g} \mathrm{g} \to \chi_{1 \c} \mathrm{g}$ & \cite{Gas87} \\ + & 89 & $\mathrm{g} \mathrm{g} \to \chi_{2 \c} \mathrm{g}$ & \cite{Gas87} \\ \hline & & d) `minimum bias' & \\ + & 91 & elastic scattering & \cite{Sch94} \\ + & 92 & single diffraction ($AB \to XB$) & \cite{Sch94} \\ + & 93 & single diffraction ($AB \to AX$) & \cite{Sch94} \\ + & 94 & double diffraction & \cite{Sch94} \\ + & 95 & low-$p_{\perp}$ production & \cite{Sjo87} \\ \hline & & e) $2 \to 1$, loop & \\ & 101 & $\mathrm{g} \mathrm{g} \to \mathrm{Z}^0$ & \\ + & 102 & $\mathrm{g} \mathrm{g} \to \H^0$ & \cite{Eic84} \\ + & 103 & $\gamma \gamma \to \H^0$ & \cite{Dre89} \\ \hline & & f) $2 \to 2$, box & \\ + & 110 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma \H^0$ & \cite{Ber85a} \\ + & 111 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{g} \H^0$ & \cite{Ell88} \\ + & 112 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_i \H^0$ & \cite{Ell88} \\ + & 113 & $\mathrm{g} \mathrm{g} \to \mathrm{g} \H^0$ & \cite{Ell88} \\ + & 114 & $\mathrm{g} \mathrm{g} \to \gamma \gamma$ & \cite{Con71,Ber84,Dic88} \\ + & 115 & $\mathrm{g} \mathrm{g} \to \mathrm{g} \gamma$ & \cite{Con71,Ber84,Dic88} \\ & 116 & $\mathrm{g} \mathrm{g} \to \gamma \mathrm{Z}^0$ & \\ & 117 & $\mathrm{g} \mathrm{g} \to \mathrm{Z}^0 \mathrm{Z}^0$ & \\ & 118 & $\mathrm{g} \mathrm{g} \to \mathrm{W}^+ \mathrm{W}^-$ & \\ & 119 & $\gamma \gamma \to \mathrm{g} \mathrm{g}$ & \\ \hline & & g) $2 \to 3$, tree & \\ + & 121 & $\mathrm{g} \mathrm{g} \to \mathrm{Q}_k \overline{\mathrm{Q}}_k \H^0$ & \cite{Kun84} \\ + & 122 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{Q}_k \overline{\mathrm{Q}}_k \H^0$ & \cite{Kun84} \\ + & 123 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_i \mathrm{f}_j \H^0$ ($\mathrm{Z} \mathrm{Z}$ fusion) & \cite{Cah84} \\ + & 124 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_k \mathrm{f}_l \H^0$ ($\mathrm{W}^+\mathrm{W}^-$ fusion) & \cite{Cah84} \\ + & 131 & $\mathrm{g} \mathrm{g} \to \mathrm{Z}^0 \mathrm{Q}_k \br{\mathrm{Q}}_k$ & \cite{Eij90} \\ \hline \end{tabular} \end{center} \end{table} \begin{table}[pt] \caption{Subprocess codes, part 4. First column is `+' for processes implemented and blank for those that are only foreseen. Second is the subprocess number \ttt{ISUB}, and third the description of the process. The final column gives references from which the cross sections have been obtained. See text for further information. \protect\label{t:procfour} } \begin{center} \begin{tabular}{|c|r|l|l|@{\protect\rule{0mm}{\tablinsep}}} \hline In & No. & Subprocess & Reference \\ \hline & & h) non-Standard Model, $2 \to 1$ & \\ + & 141 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma/\mathrm{Z}^0/\mathrm{Z}'^0$ & \cite{Alt89} \\ + & 142 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}'^+$ & \cite{Alt89} \\ + & 143 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \H^+$ & \cite{Gun87} \\ + & 144 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{R}$ & \cite{Ben85a} \\ + & 145 & $\mathrm{q}_i \ell_j \to \L_{\mathrm{Q}}$ & \cite{Wud86} \\ + & 147 & $\d \mathrm{g} \to \d^*$ & \cite{Bau90} \\ + & 148 & $\u \mathrm{g} \to \u^*$ & \cite{Bau90} \\ + & 149 & $\mathrm{g} \mathrm{g} \to \eta_{\mrm{techni}}$ & \cite{Eic84,App92} \\ + & 151 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \H'^0$ & \cite{Eic84} \\ + & 152 & $\mathrm{g} \mathrm{g} \to \H'^0$ & \cite{Eic84} \\ + & 153 & $\gamma \gamma \to \H'^0$ & \cite{Dre89} \\ + & 156 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{A}^0$ & \cite{Eic84} \\ + & 157 & $\mathrm{g} \mathrm{g} \to \mathrm{A}^0$ & \cite{Eic84} \\ + & 158 & $\gamma \gamma \to \mathrm{A}^0$ & \cite{Dre89} \\ \hline & & i) non-Standard Model, $2 \to 2$ and $2 \to 3$ & \\ + & 161 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_k \H^+$ & \cite{Bar88} \\ + & 162 & $\mathrm{q} \mathrm{g} \to \ell \L_{\mathrm{Q}}$ & \cite{Hew88} \\ + & 163 & $\mathrm{g} \mathrm{g} \to \L_{\mathrm{Q}} \br{\L}_{\mathrm{Q}}$ & \cite{Hew88,Eic84} \\ + & 164 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \L_{\mathrm{Q}} \br{\L}_{\mathrm{Q}}$ & \cite{Hew88} \\ + & 165 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{f}_k \overline{\mathrm{f}}_k$ (via $\gamma^* / \Z^0$) & \cite{Eic84,Lan91} \\ + & 166 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{f}_k \overline{\mathrm{f}}_l$ (via $\mathrm{W}^{\pm}$) & \cite{Eic84,Lan91} \\ + & 167 & $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \d^*$ & \cite{Bau90} \\ + & 168 & $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \u^*$ & \cite{Bau90} \\ + & 171 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{Z}^0 \H'^0$ & \cite{Eic84} \\ + & 172 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}^+ \H'^0$ & \cite{Eic84} \\ + & 173 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_i \mathrm{f}_j \H'^0$ ($\mathrm{Z} \mathrm{Z}$ fusion) & \cite{Cah84} \\ + & 174 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_k \mathrm{f}_l \H'^0$ ($\mathrm{W}^+\mathrm{W}^-$ fusion) & \cite{Cah84} \\ + & 176 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{Z}^0 \mathrm{A}^0$ & \cite{Eic84} \\ + & 177 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}^+ \mathrm{A}^0$ & \cite{Eic84} \\ + & 178 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_i \mathrm{f}_j \mathrm{A}^0$ ($\mathrm{Z} \mathrm{Z}$ fusion) & \cite{Cah84} \\ + & 179 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_k \mathrm{f}_l \mathrm{A}^0$ ($\mathrm{W}^+\mathrm{W}^-$ fusion) & \cite{Cah84} \\ + & 181 & $\mathrm{g} \mathrm{g} \to \mathrm{Q}_k \overline{\mathrm{Q}}_k \H'^0$ & \cite{Kun84} \\ + & 182 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{Q}_k \overline{\mathrm{Q}}_k \H'^0$ & \cite{Kun84} \\ + & 186 & $\mathrm{g} \mathrm{g} \to \mathrm{Q}_k \overline{\mathrm{Q}}_k \mathrm{A}^0$ & \cite{Kun84} \\ + & 187 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{Q}_k \overline{\mathrm{Q}}_k \mathrm{A}^0$ & \cite{Kun84} \\ \hline \end{tabular} \end{center} \end{table} In the following, $\mathrm{f}_i$ represents a fundamental fermion of flavour $i$, i.e. $\d$, $\u$, $\mathrm{s}$, $\c$, $\b$, $\t$, $\mathrm{l}$, $\mathrm{h}$, $\mathrm{e}^-$, $\nu_{\mathrm{e}}$, $\mu^-$, $\nu_{\mu}$, $\tau^-$, $\nu_{\tau}$, $\chi^-$ or $\nu_{\chi}$. A corresponding antifermion is denoted by $\overline{\mathrm{f}}_i$. In several cases, some classes of fermions are explicitly excluded, since they do not couple to the $\mathrm{g}$ or $\gamma$ (no $\e^+\e^- \to \mathrm{g} \mathrm{g}$, e.g.). When processes have only been included for quarks, while leptons might also have been possible, the notation $\mathrm{q}_i$ is used. A lepton is denoted by $\ell$; in a few cases neutrinos are also lumped under this heading. In processes where fermion masses are explicitly included in the matrix elements, an $\mathrm{F}$ is used to denote an arbitrary fermion and a $\mathrm{Q}$ a quark. Flavours appearing already in the initial state are denoted by indices $i$ and $j$, whereas new flavours in the final state are denoted by $k$ and $l$. Charge-conjugate channels are always assumed included as well (where separate), and processes involving a $\mathrm{W}^+$ also imply those involving a $\mathrm{W}^-$. Wherever $\mathrm{Z}^0$ is written, it is understood that $\gamma^*$ and $\gamma^* / \Z^0$ interference should be included as well (with possibilities to switch off either, if so desired). In some cases this is not fully implemented, see further below. Correspondingly, $\mathrm{Z}'^0$ denotes the complete set $\gamma^*/\mathrm{Z}^0/\mathrm{Z}'^0$ (or some subset of it). Thus the notation $\gamma$ is only used for a photon on the mass shell. In the last column of the tables below, references are given to works from which formulae have been taken. Sometimes these references are to the original works on the subject, sometimes only to the place where the formulae are given in the most convenient or accessible form, or where chance lead us. Apologies to all matrix-element calculators who are not mentioned. However, remember that this is not a review article on physics processes, but only a way for readers to know what is actually found in the program, for better or worse. In several instances, errata have been obtained from the authors. Often the formulae given in the literature have been generalized to include trivial radiative corrections, Breit--Wigner line shapes with $\hat{s}$-dependent widths (see section \ref{ss:kinemreson}), etc. The following sections contain some useful comments on the processes included in the program, grouped by physics interest rather than sequentially by ISUB or \ttt{MSEL} code (see \ref{ss:PYswitchkin} for further information on the \ttt{MSEL} code). The different ISUB and \ttt{MSEL} codes that can be used to simulate the different groups are given. ISUB codes within brackets indicate the kind of processes that indirectly involve the given physics topic, although only as part of a larger whole. Some obvious examples, such as the possibility to produce jets in just about any process, are not spelled out in detail. The text at times contains information on which special switches or parameters are of particular interest to a given process. All these switches are described in detail in section \ref{ss:PYswitchpar}, but are alluded to here so as to provide a more complete picture of the possibilities available for the different subprocesses. However, the list of possibilities is certainly not exhausted by the text below. \subsection{QCD Processes} In this section we discuss scatterings exclusively between coloured partons --- a process like $\e^+\e^- \to \gamma^* / \Z^0 \to \mathrm{q}\overline{\mathrm{q}}$ is also traditionally called a QCD event, but is here book-kept as $\gamma^* / \Z^0$ production. \subsubsection{QCD jets} \label{sss:QCDjetclass} \ttt{MSEL} = 1, 2 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 11 & $\mathrm{q}_i \mathrm{q}_j \to \mathrm{q}_i \mathrm{q}_j$ \\ 12 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{q}_k \overline{\mathrm{q}}_k$ \\ 13 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{g} \mathrm{g}$ \\ 28 & $\mathrm{q}_i \mathrm{g} \to \mathrm{q}_i \mathrm{g}$ \\ 53 & $\mathrm{g} \mathrm{g} \to \mathrm{q}_k \overline{\mathrm{q}}_k$ \\ 68 & $\mathrm{g} \mathrm{g} \to \mathrm{g} \mathrm{g}$ \\ \end{tabular} No higher-order processes are explicitly included, nor any higher-order loop corrections to the $2 \to 2$ processes. However, by initial- and final-state QCD radiation, multijet events are being generated, starting from the above processes. The shower rate of multijet production is clearly uncertain by some amount, especially for well-separated jets. A string-based fragmentation scheme such as the Lund model needs cross sections for the different colour flows; these have been calculated in \cite{Ben84} and differ from the usual calculations by interference terms of the order $1/N_C^2$. By default, the standard QCD expressions for the differential cross sections are used. In this case, the interference terms are distributed on the various colour flows according to the pole structure of the terms. However, the interference terms can be excluded, by changing \ttt{MSTP(34)} As an example, consider subprocess 28, $\mathrm{q} \mathrm{g} \to \mathrm{q} \mathrm{g}$. The total cross section for this process, obtained by summing and squaring the Feynman $\hat{s}$-, $\hat{t}$-, and $\hat{u}$-channel graphs, is \cite{Com77} \begin{equation} 2 \left( 1 - \frac{\hat{u}\hat{s}}{\hat{t}^2} \right) - \frac{4}{9} \left( \frac{\hat{s}}{\hat{u}} + \frac{\hat{u}}{\hat{s}} \right) - 1 ~. \end{equation} (An overall factor $\pi \alpha_{\mathrm{s}}^2/\hat{s}^2$ is ignored.) Using the identity of the Mandelstam variables for the massless case, $\hat{s} + \hat{t} + \hat{u} = 0$, this can be rewritten as \begin{equation} \frac{\hat{s}^2 + \hat{u}^2}{\hat{t}^2} - \frac{4}{9} \left( \frac{\hat{s}}{\hat{u}} + \frac{\hat{u}}{\hat{s}} \right) ~. \end{equation} On the other hand, the cross sections for the two possible colour flows of this subprocess are \cite{Ben84} \begin{eqnarray} A: & & \frac{4}{9} \left( 2 \frac{\hat{u}^2}{\hat{t}^2} - \frac{\hat{u}}{\hat{s}} \right) ~; \nonumber \\ B: & & \frac{4}{9} \left( 2 \frac{\hat{s}^2}{\hat{t}^2} - \frac{\hat{s}}{\hat{u}} \right) ~. \end{eqnarray} Colour configuration $A$ is one in which the original colour of the $\mathrm{q}$ annihilates with the anticolour of the $\mathrm{g}$, the $\mathrm{g}$ colour flows through, and a new colour--anticolour is created between the final $\mathrm{q}$ and $\mathrm{g}$. In colour configuration $B$, the gluon anticolour flows through, but the $\mathrm{q}$ and $\mathrm{g}$ colours are interchanged. Note that these two colour configurations have different kinematics dependence. For \ttt{MSTP(34)=0}, these are the cross sections actually used. The sum of the $A$ and $B$ contributions is \begin{equation} \frac{8}{9} \frac{\hat{s}^2 + \hat{u}^2}{\hat{t}^2} - \frac{4}{9} \left( \frac{\hat{s}}{\hat{u}} + \frac{\hat{u}}{\hat{s}} \right) ~. \end{equation} The difference between this expression and that of \cite{Com77}, corresponding to the interference between the two colour-flow configurations, is then \begin{equation} \frac{1}{9} \frac{\hat{s}^2 + \hat{u}^2}{\hat{t}^2} ~, \end{equation} which can be naturally divided between colour flows $A$ and $B$: \begin{eqnarray} A: & & \frac{1}{9} \frac{\hat{u}^2}{\hat{t}^2} ~; \nonumber \\ B: & & \frac{1}{9} \frac{\hat{s}^2}{\hat{t}^2} ~. \end{eqnarray} For \ttt{MSTP(34)=1}, the standard QCD matrix element is therefore used, with the same relative importance of the two colour configurations as above. Similar procedures are followed also for the other QCD subprocesses. All the matrix elements in this group are for massless quarks (although final-state quarks are of course put on the mass shell). As a consequence, cross sections are divergent for $p_{\perp} \to 0$, and some kind of regularization is required. Normally you are expected to set the desired $p_{\perp\mathrm{min}}$ value in \ttt{CKIN(3)}. The new flavour produced in the annihilation processes (ISUB = 12 and 53) is determined by the flavours allowed for gluon splitting into quark--antiquark; see switch \ttt{MDME}. \subsubsection{Heavy flavours} \label{sss:heavflavclass} \ttt{MSEL} = 4, 5, 6, 7, 8 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 81 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{Q}_k \overline{\mathrm{Q}}_k$ \\ 82 & $\mathrm{g} \mathrm{g} \to \mathrm{Q}_k \overline{\mathrm{Q}}_k$ \\ (83) & $\mathrm{q}_i \mathrm{f}_j \to Q_k f_l$ \\ \end{tabular} The matrix elements in this group differ from the corresponding ones in the group above in that they correctly take into account the quark masses. As a consequence, the cross sections are finite for $p_{\perp} \to 0$. It is therefore not necessary to introduce any special cuts. The two first processes that appear here are the dominant lowest-order graphs in hadron colliders --- a few other graphs will be mentioned later, such as process 83, which is important for a heavy top. The choice of flavour to produce is according to a hierarchy of options: \begin{Enumerate} \item if \ttt{MSEL=4-8} then the flavour is set by the \ttt{MSEL} value; \item else if \ttt{MSTP(7)=1-8} then the flavour is set by the \ttt{MSTP(7)} value; \item else the flavour is determined by the heaviest flavour allowed for gluon splitting into quark--antiquark; see switch \ttt{MDME}. \end{Enumerate} Note that only one heavy flavour is allowed at a time; if more than one is turned on in \ttt{MDME}, only the heaviest will be produced (as opposed to the case for ISUB = 12 and 53 above, where more than one flavour is allowed simultaneously). The lowest-order processes listed above just represent one source of heavy-flavour production. Heavy quarks can also be present in the parton distributions at the $Q^2$ scale of the hard interaction, leading to processes like $\mathrm{Q} \mathrm{g} \to \mathrm{Q} \mathrm{g}$, so-called flavour excitation, or they can be created by gluon splittings $\mathrm{g} \to \mathrm{Q} \overline{\mathrm{Q}}$ in initial- or final-state shower evolution. In fact, as the c.m. energy is increased, these other processes gain in importance relative to the lowest-order production graphs above. As as example, only 10\% of the $\b$ production at LHC energies come from the lowest-order graphs. The figure is even smaller for charm, while it is at or above 50\% for top. At LHC energies, the specialized treatment described in this subsection is therefore only of interest for top (and potential fourth-generation quarks) --- the higher-order corrections can here be approximated by an effective $K$ factor, except possibly in some rare corners of phase space. For charm and bottom, on the other hand, it is necessary to simulate the full event sample (within the desired kinematics cuts), and then only keep those events with $\b/\c$ either from lowest-order production, or flavour excitation, or gluon splitting. Obviously this may be a time-consuming enterprise --- although the probability for a high-$p_{\perp}$ event to contain (at least) one charm or bottom pair is fairly large, most of these heavy flavours are carrying a small fraction of the total $p_{\perp}$ flow of the jets, and therefore do not survive normal experimental cuts. As an aside, it is not only for the lowest-order graphs that events may be generated with a guaranteed heavy-flavour content. One may also generate the flavour excitation process by itself, in the massless approximation, using ISUB = 28 and setting the \ttt{KFIN} array appropriately. No trick exists to force the gluon splittings without introducing undesirable biases, however. The cross section for a heavy quark pair close to threshold can be modified according to the formulae of \cite{Fad90}, see \ttt{MSTP(35)}. Here threshold effects due to $\mathrm{Q}\overline{\mathrm{Q}}$ bound-state formation are taken into account in a smeared-out, average sense. Then the na\"{\i}ve cross section is multiplied by the squared wave function at the origin. In a colour-singlet channel this gives a net enhancement of the form \begin{equation} |\Psi^{(s)}(0)|^2 = \frac{X_{(s)}}{1 - \exp(- X_{(s)})} ~ , ~~~ \mrm{where}~ X_{(s)} = \frac{4}{3} \frac{\pi \alpha_{\mathrm{s}}}{\beta} ~, \label{pp:threshenh} \end{equation} while in a colour octet channel there is a net suppression given by \begin{equation} |\Psi^{(8)}(0)|^2 = \frac{X_{(8)}}{\exp(- X_{(8)}) -1} ~ , ~~~ \mrm{where}~ X_{(8)} = \frac{1}{6} \frac{\pi \alpha_{\mathrm{s}}}{\beta} ~. \label{pp:threshsup} \end{equation} The $\alpha_{\mathrm{s}}$ factor in this expression is related to the energy scale of bound-state formation; it is selected independently from the one of the standard production cross section. The presence of a threshold factor affects the total rate and also kinematical distributions. Heavy flavours, i.e. top and fourth generation, are assumed to be so short-lived that they decay before they have time to hadronize. This means that the light quark in the decay $\mathrm{Q} \to \mathrm{W}^{\pm} \mathrm{q}$ inherits the colour of the heavy one. The new {\tsc{Pythia}} description represents a change of philosophy compared to previous versions, formulated at a time when the top was thought to be much lighter than is believed currently. However, optionally the old description may still be used, where top hadrons are formed and these subsequently allowed to decay; see \ttt{MSTP(48)} and \ttt{MSTP(49)}. For event shapes the difference between the two time orderings normally has only marginal effects \cite{Sjo92a}. It should be noted that cross section calculations are different in the two cases. If the top (or the fourth generation fermion) is assumed short-lived, then it is treated like a resonance in the sense of section \ref{sss:resdecaycross}, i.e. the cross-section is reduced so as only to correspond to the channels left open by the user. This also includes the restrictions on secondary decays, i.e. on the decays of a $\mathrm{W}^+$ or a $\H^+$ produced in the top decay. If the top is allowed to form hadrons, no such reduction takes place. Branching ratios then have to be folded in by hand to get the correct cross sections. The logic behind this difference is that if hadronization takes place, one would be allowed e.g. to decay the $\mathrm{T}^0$ and $\mathrm{T}^+$ meson according to different branching ratios. But which $\mathrm{T}$ mesons are to be formed is not known at the top quark creation, so one could not weight for that. For a $\t$ quark which decays rapidly this ambiguity does not exist, and so a reduction factor can be introduced directly coupled to the $\t$ quark production process. This rule about cross-section calculations applies to all the processes explicitly set up to handle heavy flavour creation. In addition to the ones above, this means all the ones in Tables \ref{t:procone}--\ref{t:procfour} where the fermion final state is given as capital letters (`$\mathrm{Q}$' and `$\mathrm{F}$') and also flavours produced in resonance decays ($\mathrm{Z}^0$, $\mathrm{W}^{\pm}$, $\H^0$, etc., including processes 165 and 166). However, heavy flavours can also be produced in a process such as 31, $\mathrm{q}_i \mathrm{g} \to \mathrm{q}_k \mathrm{W}^{\pm}$, where $\mathrm{q}_k$ could be a top quark. In this case, the thrust of the description is clearly on light flavours --- the kinematics of the process is formulated in the massless fermion limit --- so any top production is purely incidental. Since here the choice of scattered flavour is only done at a later stage, the top branching ratios are not correctly folded in to the hard scattering cross section. So, for applications like these, it is not recommended to restrict the allowed top decay modes. Often one might like to get rid of the possibility of producing top together with light flavours. This can be done by switching off (i.e. setting \ttt{MDME(I,1)=0}) the `channels' $\d \to \mathrm{W}^- \t$, $\mathrm{s} \to \mathrm{W}^- \t$, $\b \to \mathrm{W}^- \t$, $\mathrm{g} \to \t\overline{\mathrm{t}}$ and $\gamma \to \t\overline{\mathrm{t}}$. Also any heavy flavours produced by parton shower evolution would not be correctly weighted into the cross section. However, currently top production is switched off in both initial (see \ttt{KFIN} array) and final (see \ttt{MSTJ(45)}) state radiation. \subsubsection{J/$\psi$} \label{sss:Jpsiclass} ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 86 & $\mathrm{g} \mathrm{g} \to \mathrm{J}/\psi \mathrm{g}$ \\ 87 & $\mathrm{g} \mathrm{g} \to \chi_{0 \c} \mathrm{g}$ \\ 88 & $\mathrm{g} \mathrm{g} \to \chi_{1 \c} \mathrm{g}$ \\ 89 & $\mathrm{g} \mathrm{g} \to \chi_{2 \c} \mathrm{g}$ \\ \end{tabular} One may distinguish three main sources of $\mathrm{J}/\psi$ production. \begin{Enumerate} \item Decays of $\mathrm{B}$ mesons and baryons. \item Parton-shower evolution, wherein a $\c$ and a $\overline{\mathrm{c}}$ quark produced in adjacent branchings (e.g. $\mathrm{g} \to \mathrm{g} \mathrm{g} \to \c \overline{\mathrm{c}} \c \overline{\mathrm{c}}$) turn out to have so small an invariant mass that the pair collapses to a single particle. \item Direct production, where a $\c$ quark loop gives a coupling between a set of gluons and a $\c\overline{\mathrm{c}}$ bound state. Higher-lying states, like the $\chi_c$ ones, may subsequently decay to $\mathrm{J}/\psi$. \end{Enumerate} In this section are given the main processes for the third source, intended for applications at hadron colliders at non-vanishing transverse momenta --- in the limit of $p_{\perp} \to 0$ it is necessary to include a number of $2 \to 1$ processes and to regularize divergences in the $2 \to 2$ graphs above. The cross sections depend on wave function values at the origin, see \ttt{PARP(38)} and \ttt{PARP(39)}. A review of the physics issues involved may be found in \cite{Glo88} (note, however, that the choice of $Q^2$ scale is different in {\tsc{Pythia}}). \subsubsection{Minimum bias} \label{sss:minbiasclass} \ttt{MSEL} = 1, 2 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 91 & elastic scattering \\ 92 & single diffraction ($AB \to XB$) \\ 93 & single diffraction ($AB \to AX$) \\ 94 & double diffraction \\ 95 & low-$p_{\perp}$ production \\ \end{tabular} These processes are briefly discussed in section \ref{ss:nonpertproc}. Currently they are mainly intended for interactions between hadrons, although one may also consider $\gamma \mathrm{p}$ interactions in the option where the incoming photon is assumed resolved, \ttt{MSTP(14)=1} or \ttt{=2}. A possible extension to $\gamma \gamma$ interactions is not yet available. Uncertainties come from a number of sources, e.g. from the parametrizations of the various cross sections and slope parameters. In diffractive scattering, the structure of the selected hadronic system may be regulated with \ttt{MSTP(101)}. No high-$p_{\perp}$ jet production in diffractive events is included so far. The subprocess 95, low-$p_{\perp}$ events, is somewhat unique in that no meaningful physical border-line to high-$p_{\perp}$ events can be defined. Even if the QCD $2 \to 2$ high-$p_{\perp}$ processes are formally switched off, some of the generated events will be classified as belonging to this group, with a $p_{\perp}$ spectrum of interactions to match the `minimum-bias' event sample. Only with the option \ttt{MSTP(82)=0} will subprocess 95 yield strictly low-$p_{\perp}$ events, events which will then probably not be compatible with any experimental data. A number of options exist for the detailed structure of low-$p_{\perp}$ events, see in particular \ttt{MSTP(81)} and \ttt{MSTP(82)}. Further details on the model(s) for minimum-bias events are found in section \ref{ss:multint}. \subsection{Electroweak Gauge Bosons} This section covers the production and/or exchange of $\gamma$, $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$ gauge bosons, singly and in pairs. The topic of longitudinal gauge-boson scattering at high energies is deferred to the Higgs section, since the presence or absence of a Higgs here makes a big difference. \subsubsection{Prompt photon production} \label{sss:promptgammaclass} \ttt{MSEL} = 10 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 14 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{g} \gamma$ \\ 18 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma \gamma$ \\ 29 & $\mathrm{q}_i \mathrm{g} \to \mathrm{q}_i \gamma$ \\ 114 & $\mathrm{g} \mathrm{g} \to \gamma \gamma$ \\ 115 & $\mathrm{g} \mathrm{g} \to \mathrm{g} \gamma$ \\ \end{tabular} In hadron colliders, processes ISUB = 14 and 29 give the main source of single-$\gamma$ production, with ISUB = 115 giving an additional contribution which, in some kinematics regions, may become important. For $\gamma$-pair production, the process ISUB = 18 is often overshadowed in importance by ISUB = 114. Another source of photons is bremsstrahlung off incoming or outgoing quarks. This has to be treated on an equal footing with QCD parton showering. For time-like parton-shower evolution, i.e. in the final-state showering and in the side branches of the initial-state showering, photon emission may be switched on or off with \ttt{MSTJ(41)}. Photon radiation off the space-like incoming quark legs is not yet included, but should be of lesser importance for production at reasonably large $p_{\perp}$ values. Radiation off an incoming electron is included in a leading-log approximation. {\bf Warning:} the cross sections for the box graphs 114 and 115 become very complicated, numerically unstable and slow when the full quark mass dependence is included. For quark masses much below the $\hat{s}$ scale, the simplified massless expressions are therefore used --- a fairly accurate approximation. However, there is another set of subtle numerical cancellations between different terms in the massive matrix elements in the region of small-angle scattering. The associated problems have not been sorted out yet. There are therefore two possible solutions. One is to use the massless formulae throughout. The program then becomes faster and numerically stable, but does not give, for example, the characteristic dip (due to destructive interference) at top threshold. This is the current default procedure, with five flavours assumed, but this number can be changed in \ttt{MSTP(38)}. The other possibility is to impose cuts on the scattering angle of the hard process, see \ttt{CKIN(27)} and \ttt{CKIN(28)}, since the numerically unstable regions are when $|\cos\hat{\theta}|$ is close to unity. It is then also necessary to change \ttt{MSTP(38)} to 0. \subsubsection{Photoproduction and $\gamma\gamma$ physics} \label{sss:photoprodclass} \ttt{MSEL} = 1, 2, 4, 5, 6, 7, 8 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 33 & $\mathrm{q}_i \gamma \to \mathrm{q}_i \mathrm{g}$ \\ 34 & $\mathrm{f}_i \gamma \to \mathrm{f}_i \gamma$ \\ 54 & $\mathrm{g} \gamma \to \mathrm{q}_k \overline{\mathrm{q}}_k$ \\ 58 & $\gamma \gamma \to \mathrm{f}_k \overline{\mathrm{f}}_k$ \\ 80 & $\mathrm{q}_i \gamma \to \mathrm{q}_k \pi^{\pm}$ \\ 84 & $\mathrm{g} \gamma \to \mathrm{Q}_k \overline{\mathrm{Q}}_k$ \\ 85 & $\gamma \gamma \to \mathrm{F}_k \overline{\mathrm{F}}_k$ \\ \end{tabular} An (almost) real photon has both a point-like component and a hadron-like one. This means that several classes of processes may be distinguished, see section \ref{sss:photoprod}. \begin{Enumerate} \item The processes listed above are possible when the photon interacts as a point-like particle, i.e. couples directly to quarks and leptons. \item When the photon acts like a hadron, i.e. is resolved in a partonic substructure, then high-$p_{\perp}$ parton--parton interactions are possible, as already described in sections \ref{sss:QCDjetclass} and \ref{sss:promptgammaclass}. These interactions may be further subdivided into VMD and anomalous ones \cite{Sch93,Sch93a}. \item A hadron-like photon can also produce the equivalent of the minimum bias processes of section \ref{sss:minbiasclass}. \end{Enumerate} For $\gamma\mathrm{p}$ events, we believe that the best description can be obtained when three separate event classes are combined, one for direct, one for VMD and one for anomalous events, see the detailed description in \cite{Sch93,Sch93a}. These correspond to \ttt{MSTP(14)} being 0, 2 and 3, respectively. The direct and anomalous components are high-$p_{\perp}$ only, while VMD contains both high-$p_{\perp}$ and low-$p_{\perp}$ events. The option \ttt{MSTP(14)=1} combines the VMD and anomalous parts of the photon into one single resolved photon concept, which therefore is less precise than the full subdivision. When combining three runs to obtain the totality of $\gamma\mathrm{p}$ interactions, to the best of our knowledge, it is necessary to choose the $p_{\perp}$ cut-offs with some care, so as to represent the expected total cross section. \begin{Itemize} \item The direct processes only depend on the \ttt{CKIN(3)} cut-off of the generation, with preferred value 0.5 GeV \cite{Sch93,Sch93a}. Since this value is so low, one must remember to reduce a few other defaults values: \ttt{CKIN(1)=2.*CKIN(3)}, \ttt{CKIN(5)=CKIN(6)=0.5*CKIN(3)}. For the same reason it is recommended to include a dampening of proton parton distributions, \ttt{MSTP(57)=2}. \item The VMD processes work as ordinary hadron--hadron ones, i.e. one obtains both low- and high-$p_{\perp}$ events by default, with dividing line set by \ttt{PARP(81)} (or \ttt{PARP(82)}, depending on minijet unitarization scheme). \item For the anomalous, finally, the minimal $p_{\perp}$ of the $\gamma \to \mathrm{q}\overline{\mathrm{q}}$ branching is set in \ttt{PARP(15)}. The default is 0.5 GeV, in agreement with the recommended cutoff for the same vertex in direct proccesses. In addition, a lower \ttt{CKIN(3)} cut-off should be selected for the hard interactions. This needs some fine-tuning, which in principle should be done separately for each c.m. energy. A good first approximation in the HERA energy range (but not beyond 300 GeV) is \ttt{CKIN(3)}$=1.50 + 0.0035 \, E_{\mrm{cm}}$. \end{Itemize} The processes in points 1 and 2 can be simulated either with a photon beam or with an electron beam. For a photon beam it is necessary to use option \ttt{MSTP(14)} to switch between a point-like and a resolved photon --- it is not possible to simulate the two sets of processes in a single run. An electron by default is assumed to contain photons, but this can be switched off by \ttt{MSTP(11)=0}. To have quark and gluon distributions inside the photon (itself inside the electron), \ttt{MSTP(12)=1} must be used. For the electron, the two kinds of processes may be generated together, unlike for the photon. It is not possible to have also the low-$p_{\perp}$ physics (including multiple interactions in high-$p_{\perp}$ events) for an electron beam. Kindly note that subprocess 34 contains both the scattering of an electron off a photon and the scattering of a quark (inside a photon inside an electron) off a photon; the former can be switched off with the help of the \ttt{KFIN} array. If you are only concerned with standard QCD physics, the option \ttt{MSTP(14)=10} gives an automatic mixture of the VMD, direct and anomalous event classes. The mixture is properly given according to the relative cross sections. Whenever possible, this option is therefore preferrable in terms of user-friendliness. However, it can only work because of a completely new layer of administration, not found anywhere else in {\tsc{Pythia}}. For instance, a subprocess like $\mathrm{q}\mathrm{g} \to \mathrm{q}\mathrm{g}$ is allowed in several of the classes, but appears with different sets of parton distributions and different $p_{\perp}$ cut-offs in each of these, so that it is necessary to switch gears between each event in the generation. It is therefore not possible to avoid a number of restrictions on what you can do in this case: \begin{Itemize} \item The \ttt{MSTP(14)=10} option can only be used for incoming photon beams, i.e. when \ttt{'gamma'} is the argument in the \ttt{PYINIT} call. A convolution with the bremsstrahlung photon spectrum in an electron beam may come one day, but not in the immediate future. \item The machinery has only been set up to generate standard QCD physics, specifically either `minimum-bias' one or high-$p_{\perp}$ jets. For minimum bias, you are not allowed to use the \ttt{CKIN} variables at all. This is not a major limitation, since it is in the spirit of minimum-bias physics not to impose any contraints on allowed jet production. (If you still do, these cuts will be ineffective for the VMD processes but take effect for the other ones, giving inconsistencies.) The minimum-bias physics option is obtained by default; by switching from \ttt{MSEL=1} to \ttt{MSEL=2} also the elastic and diffractive components of the VMD part are included. High-$p_{\perp}$ jet production is obtained by setting the \ttt{CKIN(3)} cut-off larger than each of the (energy-dependent) cut-off scales for the VMD, direct and anomalous components; typically this means at least 3 GeV. For lower input \ttt{CKIN(3)} values the program will automatically switch back to minimum-bias physics. \item Some variables are internally recalculated and reset: \ttt{CKIN(1)}, \ttt{CKIN(3)}, \ttt{CKIN(5)}, \ttt{CKIN(6)}, \ttt{MSTP(57)}, \ttt{MSTP(85)}, \ttt{PARP(2)}, \ttt{PARP(81)}, \ttt{PARP(82)}, \ttt{PARU(115)} and \ttt{MDME(22,J)}. This is because they must have values that depend on the component studied. These variables can therefore not be modified without changing \ttt{PYINPR} and recompiling the program, which obviously is a major exercise. \item Pileup events are not at all allowed. \end{Itemize} Also, a warning about the usage of \tsc{Pdflib} for photons. So long as \ttt{MSTP(14)=1}, i.e. the photon is not split up, \tsc{Pdflib} is accessed by \ttt{MSTP(56)=2} and \ttt{MSTP(55)} as the parton distribution set. However, when the VMD and anomalous pieces are split, the VMD part is based on a rescaling of pion distributions by VMD factors (except for the SaS sets, that already come with a separate VMD piece). Therefore, to access \tsc{Pdflib} for \ttt{MSTP(14)=10}, it is not correct to set \ttt{MSTP(56)=2} and a photon distribution in \ttt{MSTP(55)}. Instead, one should put \ttt{MSTP(56)=2}, \ttt{MSTP(54)=2} and a pion distribution code in \ttt{MSTP(53)}, while \ttt{MSTP(55)} has no function. The anomalous part is still based on the SaS parametrization, with \ttt{PARP(15)} as main free parameter. Currently, hadrons are not defined with any photonic content. None of the processes are therefore relevant in hadron--hadron collisions. In $\mathrm{e}\mathrm{p}$ collisions, the electron can emit an almost real photon, which may interact directly or be resolved. In $\e^+\e^-$ collisions, one may have direct, singly-resolved or doubly-resolved processes. The $\gamma\gamma$ equivalent to the $\gamma\mathrm{p}$ description involves six different event classes, see section \ref{sss:photoprod}. These classes can be obtained by setting \ttt{MSTP(14)} to 0, 2, 3, 5, 6 and 7, respectively. If one combines the VMD and anomalous parts of the parton distributions of the photon, in a more coarse description, it is enough to use the \ttt{MSTP(14)} options 0, 1 and 4. The cut-off procedures follows from the ones used for the $\gamma\mathrm{p}$ ones above. Thus the direct$\times$direct and direct$\times$VMD processes require the same cut-offs as used for direct $\gamma\mathrm{p}$ events, the VMD$\times$VMD ones the same as used for VMD $\gamma\mathrm{p}$ events, and the rest (anomalous$\times$anomalous, direct$\times$anomalous and VMD$\times$anomalous) the same as used for anomalous $\gamma\mathrm{p}$ events. As with $\gamma\mathrm{p}$ events, the option \ttt{MSTP(14)=10} gives a mixture of the six possible $\gamma\gamma$ event classes. The same complications and restrictions exist here as already listed above. For normal use the advantages should outweight the disadvantages. It is hoped to extend the formalism also to mildly virtual photons. Currently this is not done. The interaction of a highly virtual photon with a real photon is included in the deep inelastic scattering formalism below, however. Process 54 generates a mixture of quark flavours; allowed flavours are set by the gluon \ttt{MDME values}. Process 58 can generate both quark and lepton pairs, according to the \ttt{MDME} values of the photon. Processes 84 and 85 are variants of these matrix elements, with fermion masses included in the matrix elements, but where only one flavour can be generated at a time. This flavour is selected as described for processes 81 and 82 in section \ref{sss:heavflavclass}, with the exception that for process 85 the `heaviest' flavour allowed for photon splitting takes to place of the heaviest flavour allowed for gluon splitting. Since lepton KF codes come after quark ones, they are counted as being `heavier', and thus take precedence if they have been allowed. Process 80 is a higher twist one. The theory for such processes is rather shaky, so results should not be taken too literally. The messy formulae given in \cite{Bag82} have not been programmed in full, instead the pion form factor has been parametrized as $Q^2 F_{\pi}(Q^Q) \approx 0.55 / \ln Q^2$, with $Q$ in GeV. \subsubsection{Deep inelastic scattering} \label{sss:DISclass} \ttt{MSEL} = 1, 2, 35, 36, 37, 38 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 10 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_k \mathrm{f}_l$ \\ 83 & $\mathrm{q}_i \mathrm{f}_j \to \mathrm{Q}_k \mathrm{f}_l$ \\ \end{tabular} The `deep inelastic scattering' (DIS) processes, i.e. $t$-channel electroweak gauge boson exchange, are traditionally associated with interactions between a lepton or neutrino and a hadron, but processes 10 and 83 can equally well be applied for $\mathrm{q}\q$ scattering in hadron colliders (with a cross section much smaller than corresponding QCD processes, however). If applied to incoming $\e^+\e^-$ beams, process 10 corresponds to Bhabha scattering. For process 10 both $\gamma$, $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$ exchange contribute, including interference between $\gamma$ and $\mathrm{Z}^0$. The switch \ttt{MSTP(21)} may be used to restrict to only some of these, e.g. neutral or charged current only. The option \ttt{MSTP(14)=10} (see previous section) has now been extended so that it also works for deep inelastic sacattering of an electron off a (real) photon, i.e. process 10. What is obtained is a mixture of the photon acting as a vector meson and it acting as an anomalous state. This should therefore be the sum of what can be obtained with \ttt{MSTP(14)=2} and \ttt{=3}. It is distinct from \ttt{MSTP(14)=1} in that different sets are used for the parton distributions --- in \ttt{MSTP(14)=1} all the contributions to the photon distributions are lumped together, while they are split in VMD and anomalous parts for \ttt{MSTP(14)=10}. Also the beam remnant treatment is different, with a simple Gaussian distribution (at least by default) for \ttt{MSTP(14)=1} and the VMD part of \ttt{MSTP(14)=10}, but a powerlike distribution $\d k_{\perp}^2 / k_{\perp}^2$ between \ttt{PARP(15)} and $Q$ for the anomalous part of \ttt{MSTP(14)=10}. To access this option for $\mathrm{e}$ and $\gamma$ as incoming beams, it is only necessary to set \ttt{MSTP(14)=10} and keep \ttt{MSEL} at its default value. Unlike the corresponding option for $\gamma\mathrm{p}$ and $\gamma\gamma$, no cuts are overwritten, i.e. it is still the responsability of the user to set these appropriately. Cuts especially appropriate for DIS usage include either \ttt{CKIN(21)-CKIN(22)} or \ttt{CKIN(23)-CKIN(24)} for the $x$ range (former or latter depending on which side is the incoming real photon), \ttt{CKIN(35)-CKIN(36)} for the $Q^2$ range, and \ttt{CKIN(39)-CKIN(40)} for the $W^2$ range. In principle, the DIS $x$ variable of an event corresponds to the $x$ value stored in \ttt{PARI(33)} or \ttt{PARI(34)}, depending on which side the incoming hadron is on, while the DIS $Q^2 = -\hat{t} = $\ttt{-PARI(15)}. However, just like initial- and final-state radiation can shift jet momenta, they can modify the momentum of the scattered lepton. Therefore the DIS $x$ and $Q^2$ variables are not automatically conserved. An option, on by default, exists in \ttt{MSTP(23)}, where the event can be `modified back' so as to conserve $x$ and $Q^2$, but this option is still rather primitive and should not be taken too literally. Process 83 is the equivalent of process 10 for $\mathrm{W}^{\pm}$ exchange only, but with the heavy-quark mass included in the matrix element. In hadron colliders it is mainly of interest for the production of very heavy flavours, where the possibility of producing just one heavy quark is kinematically favoured over pair production. The selection of the heavy flavour is already discussed in section \ref{sss:heavflavclass}. \subsubsection{Single $\mathrm{W}/\mathrm{Z}$ production} \label{sss:WZclass} \ttt{MSEL} = 11, 12, 13, 14, 15, (21) \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 1 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma^* / \Z^0$ \\ 2 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}^+$ \\ 15 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{g} (\gamma^* / \Z^0)$ \\ 16 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{g} \mathrm{W}^+$ \\ 19 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma (\gamma^* / \Z^0)$ \\ 20 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \gamma \mathrm{W}^+$ \\ 30 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_i (\gamma^* / \Z^0)$ \\ 31 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_k \mathrm{W}^+$ \\ 35 & $\mathrm{f}_i \gamma \to \mathrm{f}_i (\gamma^* / \Z^0)$ \\ 36 & $\mathrm{f}_i \gamma \to \mathrm{f}_k \mathrm{W}^+$ \\ 131 & $\mathrm{g} \mathrm{g} \to \mathrm{Z}^0 \mathrm{Q}_k \overline{\mathrm{Q}}_k$ \\ (141) & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma/\mathrm{Z}^0/\mathrm{Z}'^0$ \\ \end{tabular} This group consists of $2 \to 1$ processes, i.e. production of a single resonance, and $2 \to 2$ processes, where the resonance is recoiling against a jet or a photon. The process 141, which also is listed here, is described further elsewhere. With initial-state showers turned on, the $2 \to 1$ processes also generate additional jets; in order to avoid double-counting, the corresponding $2 \to 2$ processes should therefore not be turned on simultaneously. The basic rule is to use the $2 \to 1$ processes for inclusive generation of $\mathrm{W}/\mathrm{Z}$, i.e. where the bulk of the events studied have $p_{\perp} \ll m_{\mathrm{W}/\mathrm{Z}}$, which is where parton showers may be expected to do a good job. For dedicated studies of $\mathrm{W}/\mathrm{Z}$ production at larger transverse momenta, the parton showers tend to underestimate the event rates. It is here better to start from the $2 \to 2$ matrix elements and add showers to these. However, the $2 \to 2$ matrix elements are divergent for $p_{\perp} \to 0$, and should not be used down to the low-$p_{\perp}$ region, or one may get unphysical cross sections. The problem of double-counting applies not only to $\mathrm{W}/\mathrm{Z}$ production in hadron colliders, but also to a process like $\e^+\e^- \to \mathrm{Z}^0 \gamma$, which clearly is part of the initial-state radiation corrections to $\e^+\e^- \to \mathrm{Z}^0$ obtained for \ttt{MSTP(11)=1}. As is the case for $\mathrm{Z}$ production in association with jets, the $2 \to 2$ process should therefore only be used for the high-$p_{\perp}$ region. The $\mathrm{Z}^0$ of subprocess 1 includes the full interference structure $\gamma^* / \Z^0$; via \ttt{MSTP(43)} you can select to produce only $\gamma^*$, only $\mathrm{Z}^0$, or the full $\gamma^* / \Z^0$. The same holds true for the $\mathrm{Z}'^0$ of subprocess 141; via \ttt{MSTP(44)} any combination of $\gamma^*$, $\mathrm{Z}^0$ and $\mathrm{Z}'^0$ can be selected. Thus, subprocess 141 with \ttt{MSTP(44)=4} is essentially equivalent to subprocess 1 with \ttt{MSTP(43)=3}; however, process 141 also includes the possibility of a decay into Higgses. Also processes 15, 19, 30 and 35 contain the full mixture of $\gamma^* / \Z^0$, with \ttt{MSTP(43)} available to change this. Only the $\mathrm{Z}^0$ that appears in process 131 does not contain the $\gamma^*$ contribution. Note that process 1, with only $\mathrm{q}\overline{\mathrm{q}} \to \gamma^* \to \ell^+ \ell^-$ allowed, and studied in the region well below the $\mathrm{Z}^0$ mass, is what is conventionally called Drell--Yan. This latter process therefore does not appear under a separate heading, but can be obtained by a suitable setting of switches and parameters. A process like $\mathrm{f}_i \overline{\mathrm{f}}_j \to \gamma \mathrm{W}^+$ is only included in the limit that the $\gamma$ is emitted in the `initial state', while the possibility of a final-state radiation off the $\mathrm{W}^+$ decay products is not explicitly included (but can be obtained implicitly by the parton-shower machinery) and various interference terms are not at all present. Some caution must therefore be exercised; see also section \ref{sss:WZpairclass} for related comments. For the $2 \to 1$ processes, the Breit--Wigner includes an $\hat{s}$-dependent width, which should provide an improved description of line shapes. In fact, from a line-shape point of view, process 1 should provide a more accurate simulation of $\e^+\e^-$ annihilation events than the dedicated $\e^+\e^-$ generation scheme of {\tsc{Jetset}} (see section \ref{ss:eematrix}). However, the $p_{\perp}$ distribution of radiated initial-state photons is probably still better modelled in the {\tsc{Jetset}} routines. Another difference is that {\tsc{Jetset}} only allows the generation of $\gamma^* / \Z^0 \to \mathrm{q}\overline{\mathrm{q}}$, while process 1 additionally contains $\gamma^* / \Z^0 \to \ell^+ \ell^-$ and $\gamma^* / \Z^0 \to \nu \br{\nu}$. The parton-shower and fragmentation descriptions are the same, but the {\tsc{Pythia}} implementation has not been interfaced with the first- and second-order matrix-element options available in {\tsc{Jetset}}. Almost all processes in this group have been included with the correct angular distribution in the subsequent $\mathrm{W}/\mathrm{Z} \to \mathrm{f} \overline{\mathrm{f}}$ decays. The exception is process 36, where currently the $\mathrm{W}$ decays isotropically. The process $\e^+\e^- \to \mathrm{e}^+ \mathrm{e}^- \mathrm{Z}^0$ can be simulated in two different ways. One is to make use of the $\mathrm{e}$ `sea' distribution inside $\mathrm{e}$, i.e. have splittings $\mathrm{e} \to \gamma \to \mathrm{e}$. This can be obtained, together with ordinary $\mathrm{Z}^0$ production, by using subprocess 1, with \ttt{MSTP(11)=1} and \ttt{MSTP(12)=1}. Then the contribution of the type above is 5.0 pb for a 500 GeV $\e^+\e^-$ collider, compared with the correct 6.2 pb \cite{Hag91}. Alternatively one may use process 35, with \ttt{MSTP(11)=1} and \ttt{MSTP(12)=0}. To catch the singularity in the forward direction, regularized by the electron mass, it is necessary to set \ttt{CKIN(3)=CKIN(5)=0.01} --- using lower values will only slow down execution, not significantly increasing the cross section. One then obtains 5.1 pb, i.e. again 20\% below the correct value, but now also generates a $p_{\perp}$ distribution for the $\mathrm{Z}^0$; this is therefore to be preferred. Process 36, $\mathrm{f} \gamma \to \mathrm{f}' \mathrm{W}^{\pm}$ may have corresponding problems; except that in $\e^+\e^-$ the forward scattering amplitude for $\mathrm{e} \gamma \to \nu \mathrm{W}$ is killed (radiation zero), which means that the differential cross section is vanishing for $p_{\perp} \to 0$. It is therefore feasible to use the default \ttt{CKIN(3)} and \ttt{CKIN(5)} values in $\e^+\e^-$, and one also comes closer to the correct cross section. One single true $2 \to 3$ process is included in this class as well; namely $\mathrm{g} \mathrm{g} \to \mathrm{Z}^0 \mathrm{Q} \overline{\mathrm{Q}}$, with full massive matrix elements. The more complicated phase space and the lengthy matrix-element evaluations make this process extremely slow. With the quark flavour picked to be $\b$, it may form an important background to intermediate mass Higgs searches in the multilepton channel. The quark flavour is stored in \ttt{KFPR(131,2)}; the default is $5 = \b$. The kinematics is set up in terms of a $\mathrm{Z}^0$ recoiling against the $\mathrm{Q} \overline{\mathrm{Q}}$ system, and all ordinary kinematics cut for a $2 \to 2$ process can be used on this level, including \ttt{CKIN(43)} and \ttt{CKIN(44)} to restrict the range of the $\mathrm{Q} \overline{\mathrm{Q}}$ invariant mass. In addition, for this process alone, \ttt{CKIN(51) - CKIN(54)} can be used to set the $p_{\perp}$ range of the two quarks; as is to be expected, that of the $\mathrm{Z}^0$ is set by \ttt{CKIN(3) - CKIN(4)}. Since the optimization procedure is not set up to probe the full multidimensional phase space allowed in this process, maximum violations may be quite large. It may then be useful to make a preliminary run to find how big the violations are in total, and then use the \ttt{MSTP(121)=1} option in the full run. \subsubsection{$\mathrm{W}/\mathrm{Z}$ pair production} \label{sss:WZpairclass} \ttt{MSEL} = 15 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 22 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to (\gamma^* / \Z^0) (\gamma^* / \Z^0)$ \\ 23 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{Z}^0 \mathrm{W}^+$ \\ 25 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{W}^+ \mathrm{W}^-$ \\ 69 & $\gamma \gamma \to \mathrm{W}^+ \mathrm{W}^-$ \\ 70 & $\gamma \mathrm{W}^+ \to \mathrm{Z}^0 \mathrm{W}^+$ \\ \end{tabular} In this section we mainly consider the production of $\mathrm{W}/\mathrm{Z}$ pairs by fermion--antifermion annihilation, but also include two processes which involve $\gamma/\mathrm{W}$ beams. Scatterings between gauge-boson pairs, i.e. processes like $\mathrm{W}^+ \mathrm{W}^- \to \mathrm{Z}^0 \mathrm{Z}^0$, depend so crucially on the assumed Higgs scenario that they are considered separately in section \ref{sss:heavySMHclass}. The cross sections used for the above processes are those derived in the narrow-width limit, but have been extended to include Breit--Wigner shapes with mass-dependent widths. However, one should realize that other graphs, not included here, can contribute in regions away from the $\mathrm{W}/\mathrm{Z}$ mass. This problem is especially important if several flavours coincide in the four-fermion final state. Consider, as an example, $\e^+\e^- \to \mu^+ \mu^- \nu_{\mu} \br{\nu}_{\mu}$. Not only would such a final state receive contributions from intermediate $\mathrm{Z}^0\mathrm{Z}^0$ and $\mathrm{W}^+\mathrm{W}^-$ states, but also from processes $\e^+\e^- \to \mathrm{Z}^0 \to \mu^+ \mu^-$, followed either by $\mu^+ \to \mu^+ \mathrm{Z}^0 \to \mu^+ \nu_{\mu} \br{\nu}_{\mu}$, or by $\mu^+ \to \br{\nu}_{\mu} \mathrm{W}^+ \to \br{\nu}_{\mu} \mu^+ \nu_{\mu}$. In addition, all possible interferences should be considered. Since this is not done, the processes have to be used with some sound judgement. Very often, one may wish to constrain a lepton pair mass to be close to $m_{\mathrm{Z}}$, in which case a number of the possible `other' processes are negligible. Of the above processes, the first contains the full $\mathrm{f}_i \overline{\mathrm{f}}_i \to (\gamma^* / \Z^0)(\gamma^* / \Z^0)$ structure, obtained by a straightforward generalization of the formulae in ref. \cite{Gun86} (done by the present author). Of course, the possibility of there being significant contributions from graphs that are not included is increased, in particular if one $\gamma^*$ is very light and therefore could be a bremsstrahlung-type photon. It is possible to use \ttt{MSTP(43)} to recover the pure $\mathrm{Z}^0$ case, i.e. $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{Z}^0 \mathrm{Z}^0$ exclusively. In processes 23 and 70, only the pure $\mathrm{Z}^0$ contribution is included. Full angular correlations are included for the first three processes, i.e. the full $2 \to 2 \to 4$ matrix elements are included in the resonance decays, including the appropriate $\gamma^* / \Z^0$ interference in process 22. In the latter two processes no spin information is currently preserved, i.e. the $\mathrm{W}/\mathrm{Z}$ bosons are allowed to decay isotropically. We remind you that the mass ranges of the two resonances may be set with the \ttt{CKIN(41) - CKIN(44)} parameters; this is particularly convenient, for instance, to pick one resonance almost on the mass shell and the other not. \subsection{Higgs Production} A fair fraction of all the processes in {\tsc{Pythia}} deal with Higgs production in one form or another. This multiplication is caused by the need to consider production by several different processes, depending on Higgs mass and machine type. Further, the program contains a full two-Higgs-multiplet scenario, as predicted for example in the Minimal Supersymmetric extension of the Standard Model (MSSM). Therefore the continued discussion is, somewhat arbitrarily, subdivided into a few different scenarios. \subsubsection{Light Standard Model Higgs} \label{sss:lightSMHclass} \ttt{MSEL} = 16, 17, 18 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 3 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \H^0$ \\ 24 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{Z}^0 \H^0$ \\ 26 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}^+ \H^0$ \\ 102 & $\mathrm{g} \mathrm{g} \to \H^0$ \\ 103 & $\gamma \gamma \to \H^0$ \\ 110 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma \H^0$ \\ 111 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{g} \H^0$ \\ 112 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_i \H^0$ \\ 113 & $\mathrm{g} \mathrm{g} \to \mathrm{g} \H^0$ \\ 121 & $\mathrm{g} \mathrm{g} \to \mathrm{Q}_k \overline{\mathrm{Q}}_k \H^0$ \\ 122 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{Q}_k \overline{\mathrm{Q}}_k \H^0$ \\ 123 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_i \mathrm{f}_j \H^0$ ($\mathrm{Z}^0 \mathrm{Z}^0$ fusion) \\ 124 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_k \mathrm{f}_l \H^0$ ($\mathrm{W}^+ \mathrm{W}^-$ fusion) \\ \end{tabular} In this section we discuss the production of a reasonably light Standard Model Higgs, below 700 GeV, say, so that the narrow width approximation can be used with some confidence. Below 400 GeV there would certainly be no trouble, while above that the narrow width approximation is gradually starting to break down. In a hadron collider, the main production processes are 102, 123 and 124, i.e. $\mathrm{g}\g$, $\mathrm{Z}^0 \mathrm{Z}^0$ and $\mathrm{W}^+ \mathrm{W}^-$ fusion. In the latter two processes, it is also necessary to take into account the emission of the space-like $\mathrm{W}/\mathrm{Z}$ bosons off quarks, which in total gives the $2 \to 3$ processes above. Further processes of lower cross sections may be of interest because of easier signals. For instance, processes 24 and 26 give associated production of a $\mathrm{Z}$ or a $\mathrm{W}$ together with the $\H^0$. There is also the processes 3, 121 and 122, which involve production of heavy flavours. Process 3 contains contributions from all flavours, but is completely dominated by the subprocess $\t \overline{\mathrm{t}} \to \H^0$, i.e. by the contribution from the top sea distributions. This process is by now known to overestimate the cross section for Higgs production as compared with a more careful calculation based on the subprocess $\mathrm{g} \mathrm{g} \to \t \overline{\mathrm{t}} \H^0$ (121). The difference between the two is that in process 3 the $\t$ and $\overline{\mathrm{t}}$ are added by the initial-state shower, while in 121 the full matrix element is used. The price to be paid is that the complicated multibody phase space in process 121 makes the program run slower than with most other processes. One should therefore think twice before using it. As usual, it would be double-counting to include both 3 and 121. Process 122 is similar in structure to 121, but is less important. In both process 121 and 122 the produced quark is assumed to be a $\t$; this can be changed in \ttt{KFPR(121,2)} and \ttt{KFPR(122,2)} before initialization, however. A subprocess like 113, with a Higgs recoiling against a gluon jet, is also effectively generated by initial-state corrections to subprocess 102; thus, in order to avoid double-counting, just as for the case of $\mathrm{Z}^0/\mathrm{W}^+$ production, section \ref{sss:WZclass}, these subprocesses should not be switched on simultaneously. Process 102 should be used for inclusive production of Higgs, and 111--113 for the study of the Higgs subsample with high transverse momentum. In $\e^+\e^-$ annihilation, associated production of an $\H^0$ with a $\mathrm{Z}^0$, process 24, is usually the dominant one close to threshold, while the $\mathrm{Z}^0 \mathrm{Z}^0$ and $\mathrm{W}^+ \mathrm{W}^-$ fusion processes 123 and 124 win out at high energies. Process 103, $\gamma\gamma$ fusion, may also be of interest, in particular when the possibilities of beamstrahlung photons and backscattered photons are included. Process 110, which gives an $\H^0$ in association with a $\gamma$, is a loop process and is therefore suppressed in rate. Only for a rather massive $\H^0$ (mass above 60 GeV at LEP 1) can it start to compete with the associated production of a $\mathrm{Z}^0$, since phase space suppression is less severe for the former than for the latter. The branching ratios of the Higgs are very strongly dependent on the mass. In principle, the program is set up to calculate these correctly, as a function of the actual Higgs mass, i.e. not just at the nominal mass. However, higher-order corrections may at times be important and not fully unambiguous; see for instance \ttt{MSTP(37)}. Since the Higgs is a spin-0 particle it decays isotropically. In decay processes such as $\H^0 \to \mathrm{W}^+ \mathrm{W}^- \to 4$ fermions angular correlations are included. Also in processes 24 and 26, $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$ decay angular distributions are correctly taken into account. \subsubsection{Heavy Standard Model Higgs} \label{sss:heavySMHclass} ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 5 & $\mathrm{Z}^0 \mathrm{Z}^0 \to \H^0$ \\ 8 & $\mathrm{W}^+ \mathrm{W}^- \to \H^0$ \\ 71 & $\mathrm{Z}^0 \mathrm{Z}^0 \to \mathrm{Z}^0 \mathrm{Z}^0$ (longitudinal) \\ 72 & $\mathrm{Z}^0 \mathrm{Z}^0 \to \mathrm{W}^+ \mathrm{W}^-$ (longitudinal) \\ 73 & $\mathrm{Z}^0 \mathrm{W}^+ \to \mathrm{Z}^0 \mathrm{W}^+$ (longitudinal) \\ 76 & $\mathrm{W}^+ \mathrm{W}^- \to \mathrm{Z}^0 \mathrm{Z}^0$ (longitudinal) \\ 77 & $\mathrm{W}^+ \mathrm{W}^{\pm} \to \mathrm{W}^+ \mathrm{W}^{\pm}$ (longitudinal) \\ \end{tabular} Processes 5 and 8 are the simple $2 \to 1$ versions of what is now available in 123 and 124 with the full $2 \to 3$ kinematics. For low Higgs masses processes 5 and 8 overestimate the correct cross sections and should not be used, whereas good agreement between the $2 \to 1$ and $2 \to 3$ descriptions is observed when heavy Higgs production is studied. The subprocesses 5 and 8, $V V \to \H^0$, which contribute to the processes $V V \to V' V'$, show a bad high-energy behaviour. Here $V$ denotes a longitudinal intermediate gauge boson, $\mathrm{Z}^0$ or $\mathrm{W}^{\pm}$. This can be cured only by the inclusion of all $V V \to V' V'$ graphs, as is done in subprocesses 71, 72, 73, 76 and 77. In particular, subprocesses 5 and 8 give rise to a fictitious high-mass tail of the Higgs. If this tail is thrown away, however, the agreement between the $s$-channel graphs only (subprocesses 5 and 8) and the full set of graphs (subprocesses 71 etc.) is very good: for a Higgs of nominal mass 300 (800) GeV, a cut at 600 (1200) GeV retains 95\% (84\%) of the total cross section, and differs from the exact calculation, cut at the same values, by only 2\% (11\%) (numbers for SSC energies). With this prescription there is therefore no need to use subprocesses 71 etc. rather than subprocesses 5 and 8. For subprocess 77, there is an option, see \ttt{MSTP(45)}, to select the charge combination of the scattering $\mathrm{W}$'s: like-sign, opposite-sign (relevant for Higgs), or both. Process 77 contains a divergence for $p_{\perp} \to 0$ due to $\gamma$-exchange contributions. This leads to an infinite total cross section, which is entirely fictitious, since the simple parton-distribution function approach to the longitudinal $\mathrm{W}$ flux is not appropriate in this limit. For this process, it is therefore necessary to make use of a cut, e.g. $p_{\perp} > m_{\mathrm{W}}$. For subprocesses 71, 72, 76 and 77, an option is included (see \ttt{MSTP(46)}) whereby the user can select only the $s$-channel Higgs graph; this will then be essentially equivalent to running subprocess 5 or 8 with the proper decay channels (i.e. $\mathrm{Z}^0\mathrm{Z}^0$ or $\mathrm{W}^+\mathrm{W}^-$) set via \ttt{MDME}. The difference is that the Breit--Wigners in subprocesses 5 and 8 contain a mass-dependent width, whereas the width in subprocesses 71--77 is calculated at the nominal Higgs mass; also, higher-order corrections to the widths are treated more accurately in subprocesses 5 and 8. Further, processes 71--77 assume the incoming $\mathrm{W}/\mathrm{Z}$ to be on the mass shell, with associated kinematics factors, while processes 5 and 8 have $\mathrm{W}/\mathrm{Z}$ correctly space-like. All this leads to differences in the cross sections by up to a factor of 1.5. In the absence of a Higgs, the sector of longitudinal $\mathrm{Z}$ and $\mathrm{W}$ scattering will become strongly interacting at energies above 1 TeV. The models proposed by Dobado, Herrero and Terron \cite{Dob91} to describe this kind of physics have been included as alternative matrix elements for subprocesses 71, 72, 73, 76 and 77, selectable by \ttt{MSTP(46)}. From the point of view of the general classification scheme for subprocesses, this kind of models should appropriately be included as separate subprocesses with numbers above 100, but the current solution allows a more efficient reuse of existing code. By a proper choice of parameters, it is also here possible to simulate the production of a techni-$\rho$. Currently, the scattering of transverse gauge bosons has not been included, neither that of mixed transverse--longitudinal scatterings. These are expected to be less important at high energies, and do not contain an $\H^0$ resonance peak, but need not be entirely negligible in magnitude. As a rule of thumb, processes 71--77 should not be used for $VV$ invariant masses below 500 GeV. The decay products of the longitudinal gauge bosons are correctly distributed in angle. \subsubsection{Extended neutral Higgs sector} \label{sss:extneutHclass} \ttt{MSEL} = 19 \\ ISUB = \\ \begin{tabular}{rrrl@{\protect\rule{0mm}{\tablinsep}}} $\H^0$ & $\H'^0$ & $\mathrm{A}^0$ & \\ 3 & 151 & 156 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to X$ \\ 102 & 152 & 157 & $\mathrm{g} \mathrm{g} \to X$ \\ 103 & 153 & 158 & $\gamma \gamma \to X$ \\ 24 & 171 & 176 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{Z}^0 X$ \\ 26 & 172 & 177 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}^+ X$ \\ 123 & 173 & 178 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_i \mathrm{f}_j X$ ($\mathrm{Z} \mathrm{Z}$ fusion) \\ 124 & 174 & 179 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_k \mathrm{f}_l X$ ($\mathrm{W}^+\mathrm{W}^-$ fusion) \\ 121 & 181 & 186 & $\mathrm{g} \mathrm{g} \to \mathrm{Q}_k \overline{\mathrm{Q}}_k X$ \\ 122 & 182 & 187 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{Q}_k \overline{\mathrm{Q}}_k X$ \\ \end{tabular} \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} (141) & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma/\mathrm{Z}^0/\mathrm{Z}'^0$ \\ \end{tabular} In {\tsc{Pythia}}, the particle content of a two-Higgs-doublet scenario is included: two neutral scalar particles, 25 and 35, one pseudoscalar one, 36, and a charged doublet, $\pm 37$. (Of course, these particles may also be associated with corresponding Higgs states in larger multiplets.) By convention, we choose to call the lighter scalar Higgs $\H^0$ and the heavier $\H'^0$ --- this differs from the convention in the MSSM, where the lighter is called $\mathrm{h}^0$ and the heavier $\H^0$, but allows us to call the Higgs of the one-Higgs scenario $\H^0$. The pseudoscalar is called $\mathrm{A}^0$ and the charged $\H^{\pm}$. Charged-Higgs production is covered in section \ref{sss:chHclass}. A number of $\H^0$ processes have been duplicated for $\H'^0$ and $\mathrm{A}^0$. The correspondence between ISUB numbers is shown in the table above: the first column of ISUB numbers corresponds to $X = \H^0$, the second to $X = \H'^0$, and the third to $X = \mathrm{A}^0$. Note that several of these processes are not expected to take place at all, owing to vanishing Born term couplings. We have still included them for flexibility in simulating arbitrary couplings at the Born or loop level. A few Standard Model Higgs processes have no correspondence in the scheme above. These include \begin{Itemize} \item 5 and 8, which anyway have been superseded by 123 and 124; \item 71, 72, 73, 76 and 77, which deal with what happens if there is no light Higgs, and so is a scenario complementary to the one above, where several light Higgses are assumed; \item 110, which is mainly of interest in Standard Model Higgs searches; and \item 111, 112 and 113, which describe the high-$p_{\perp}$ tail of the Higgs production, and are less interesting for most Higgs studies. \end{Itemize} In processes 121, 122, 181, 182, 186 and 187 the recoiling heavy flavour is assumed to be top, which is the only one of interest in the Standard Model, and the one where the parton-distribution-function approach invoked in processes 3, 151 and 156 is least reliable. However, it is possible to change the quark flavour in 121 etc.; for each process ISUB this flavour is given by \ttt{KFPR(ISUB,2)}. This may become relevant if couplings to $\b\overline{\mathrm{b}}$ states are enhanced, e.g. if $\tan\beta \gg 1$ in the MSSM. By default, the $\H^0$ has the couplings of the Standard Model Higgs, while the $\H'^0$ and $\mathrm{A}^0$ have couplings set in \ttt{PARU(171) - PARU(178)} and \ttt{PARU(181) - PARU(190)}, respectively. The default values for the $\H'^0$ and $\mathrm{A}^0$ have no deep physics motivation, but are set just so that the program will not crash due to the absence of any couplings whatsoever. You should therefore set the above couplings to your desired values if you want to simulate either $\H'^0$ or $\mathrm{A}^0$. Also the couplings of the $\H^0$ particle can be modified, in \ttt{PARU(161) - PARU(165)}, provided that \ttt{MSTP(4)} is set to 1. For \ttt{MSTP(4)=2}, the mass of the $\H^0$ (in \ttt{PMAS(25,1)}) and the $\tan\beta$ value (in \ttt{PARU(141)}) are used to derive the masses of the other Higgses, as well as all Higgs couplings. \ttt{PMAS(35,1) - PMAS(37,1)} and \ttt{PARU(161) - PARU(195)} are overwritten accordingly. The relations used are the ones of the Born-level MSSM \cite{Gun90}. Today, loop corrections to those expressions have been calculated, and are known to have non-negligible effects on the resulting phenomenology. Eventually the modified relations will be included as an additional option, but this has not yet been done. Note that not all combinations of $m_{\H}$ and $\tan\beta$ are allowed; the requirement of a finite $\mathrm{A}^0$ mass imposes the constraint \begin{equation} m_{\H} < m_{\mathrm{Z}} \, \frac{\tan^2\beta - 1}{\tan^2\beta + 1}, \end{equation} or, equivalently, \begin{equation} \tan^2\beta > \frac{m_{\mathrm{Z}} + m_{\H}}{m_{\mathrm{Z}} - m_{\H}}. \end{equation} If this condition is not fulfilled, the program will crash. Process 141 can also be used to simulate $\mathrm{Z}^0 \to \H^0 \mathrm{A}^0$ and $\mathrm{Z}^0 \to \H'^0 \mathrm{A}^0$ for associated neutral Higgs production. The fact that we here make use of the $\mathrm{Z}'^0$ can easily be discounted, either by letting the relevant couplings vanish, or by the option \ttt{MSTP(44)=4}. Finally, heavier Higgses may decay into lighter ones, if kinematically allowed, in processes like $\mathrm{A}^0 \to \mathrm{Z}^0 \H^0$ or $\H^+ \to \mathrm{W}^+ \H^0$. Such modes are included as part of the general mixture of decay channels, but they can be enhanced if the uninteresting channels are switched off. \subsubsection{Charged Higgs sector} \label{sss:chHclass} \ttt{MSEL} = 23 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 143 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \H^+$ \\ 161 & $\mathrm{f}_i \mathrm{g} \to \mathrm{f}_k \H^+$ \\ (141) & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma/\mathrm{Z}^0/\mathrm{Z}'^0$ \\ \end{tabular} A charged Higgs doublet, $\H^{\pm}$, is included in the program. This doublet may be the one predicted in the MSSM scenario, see section \ref{sss:extneutHclass}, or in any other scenario. The $\tan\beta$ parameter, which is relevant also for charged Higgs couplings, is set via \ttt{PARU(141)}. The basic subprocess for charged Higgs production in hadron colliders is ISUB = 143. However, this process is dominated by $\t \overline{\mathrm{b}} \to \H^+$, and so depends on the choice of $\t$ parton distribution. A better representation is provided by subprocess 161, $\mathrm{f} \mathrm{g} \to \mathrm{f}' \H^+$; i.e. actually $\overline{\mathrm{b}} \mathrm{g} \to \overline{\mathrm{t}} \H^+$. It is therefore recommended to use 161 and not 143; to use both would be double-counting. In subprocess 141, the decay $\gamma^*/\mathrm{Z}^0/\mathrm{Z}'^0 \to \H^+ \H^-$ allows the production of a pair of charged Higgs particles. This process is especially important in $\e^+\e^-$ colliders. The coupling of the $\gamma^*$ to $\H^+ \H^-$ is determined by the charge alone, while the $\mathrm{Z}^0$ coupling is regulated by \ttt{PARU(142)}, and that of the $\mathrm{Z}'^0$ by \ttt{PARU(143)}. The $\mathrm{Z}'^0$ piece can be switched off, e.g. by \ttt{MSTP(44)=4}. An ordinary $\mathrm{Z}^0$, i.e. particle code 23, cannot be made to decay into $\H^+ \H^-$, however. A major potential source of charged Higgs production is top decay. When the top is treated as a resonance (the default option), it is possible to switch on the decay channel $\t \to \b \H^+$. Top will then decay to $\H^+$ a fraction of the time, whichever way it is produced. The branching ratio is automatically calculated, based on the $\tan\beta$ value and masses. It is possible to only have the $\H^+$ decay mode switched on, in which case the cross section is reduced accordingly. If one instead assumes that top hadrons are formed, branching ratios are not automatically calculated. However, you can set, for the generic top hadron 86, the branching ratios for the two main channels $\t \to \b \H^+$ and $\t \to \b \mathrm{W}^+$. In this option the cross section for top production will not be reduced if only the $\t \to \b \H^+$ decay is switched on, cf. section \ref{sss:resdecaycross}. \subsection{Non-Standard Physics} The number of possible non-Standard Model scenarios is essentially infinite, but many of the studied scenarios still share a lot of aspects. For instance, new $\mathrm{W}'$ and $\mathrm{Z}'$ gauge bosons can arise in a number of different ways. Therefore it still makes sense to try to cover a few basic classes of particles, with enough freedom in couplings that many kinds of detailed scenarios can be accommodated by suitable parameter choices. We have already seen one example of this, in the extended Higgs sector above. In this section a few other kinds of non-standard generic physics is discussed. Clearly many others could have been included, but there is probably only one glaring omission: currently no supersymmetric particle production has been included. One main reason for this is the large number of particles, processes, possible mass hierarchies and decay chains. \subsubsection{Fourth-generation fermions} \ttt{MSEL} = 7, 8, 37, 38 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 1 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma^* / \Z^0$ \\ 2 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}^+$ \\ 81 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{Q}_k \overline{\mathrm{Q}}_k$ \\ 82 & $\mathrm{g} \mathrm{g} \to \mathrm{Q}_k \overline{\mathrm{Q}}_k$ \\ 83 & $\mathrm{q}_i \mathrm{f}_j \to \mathrm{Q}_k \mathrm{f}_l$ \\ 84 & $\mathrm{g} \gamma \to \mathrm{Q}_k \overline{\mathrm{Q}}_k$ \\ 85 & $\gamma \gamma \to \mathrm{F}_k \overline{\mathrm{F}}_k$ \\ 141 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma/\mathrm{Z}^0/\mathrm{Z}'^0$ \\ 142 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}'^+$ \\ \end{tabular} The prospects of a fourth generation currently seem rather dim, but the appropriate flavour content is still found in the program. In fact, the fourth generation is included on an equal basis with the first three, provided \ttt{MSTP(1)=4}. Also processes other than the ones above can therefore be used, e.g. all other processes with gauge bosons, including non-standard ones such as the $\mathrm{Z}'^0$. We therefore do not repeat the descriptions found elsewhere, e.g. how to set only the desired flavour in processes 81--85. Note that it may be convenient to set \ttt{CKIN(1)} and other cuts such that the mass of produced gauge bosons is enough for the wanted particle production --- in principle the program will cope even without that, but possibly at the expense of very slow execution. \subsubsection{New gauge bosons} \ttt{MSEL} = 21, 22, 24 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 141 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \gamma/\mathrm{Z}^0/\mathrm{Z}'^0$ \\ 142 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{W}'^+$ \\ 144 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{R}$ \\ \end{tabular} The $\mathrm{Z}'^0$ of subprocess 141 contains the full $\gamma^*/\mathrm{Z}^0/\mathrm{Z}'^0$ interference structure for couplings to fermion pairs. With \ttt{MSTP(44)} it is possible to pick only a subset, e.g. only the pure $\mathrm{Z}'^0$ piece. The couplings of the $\mathrm{Z}'^0$ to quarks and leptons can be set via \ttt{PARU(121) - PARU(128)}. The eight numbers correspond to the vector and axial couplings of down-type quarks, up-type quarks, leptons and neutrinos, respectively. The default corresponds to the same couplings as that of the Standard Model $\mathrm{Z}^0$, with axial couplings $a_{\mathrm{f}} = \pm 1$ and vector couplings $v_{\mathrm{f}} = a_{\mathrm{f}} - 4 e_{\mathrm{f}} \sin^2 \! \theta_W$. This implies a resonance width that increases linearly with the mass. By a suitable choice of the parameters, it is possible to simulate just about any imaginable $\mathrm{Z}'^0$ scenario, with full interference effects in cross sections and decay angular distributions. The coupling to the decay channel $\mathrm{Z}'^0 \to \mathrm{W}^+ \mathrm{W}^-$ is regulated by \ttt{PARU(129) - PARU(130)}. The former gives the strength of the coupling, which determines the rate. The default, \ttt{PARU(129)=1.}, corresponds to the `extended gauge model' of \cite{Alt89}, wherein the $\mathrm{Z}^0 \to \mathrm{W}^+ \mathrm{W}^-$ coupling is used, scaled down by a factor $m_{\mathrm{W}}^2/m_{\mathrm{Z}'}^2$, to give a $\mathrm{Z}'^0$ partial width into this channel that again increases linearly. If this factor is cancelled, by having \ttt{PARU(129)} proportional to $m_{\mathrm{Z}'}^2/m_{\mathrm{W}}^2$, one obtains a partial width that goes like the fifth power of the $\mathrm{Z}'^0$ mass, the `reference model' of \cite{Alt89}. In the decay angular distribution one could imagine a much richer structure than is given by the one parameter \ttt{PARU(130)}. Other decay modes include $\mathrm{Z}'^0 \to \mathrm{Z}^0 \H^0$, predicted in left--right symmetric models (see \ttt{PARU(145)} and ref. \cite{Coc91}), and a number of other Higgs decay channels, see sections \ref{sss:extneutHclass} and \ref{sss:chHclass}. The $\mathrm{W}'^{\pm}$ of subprocess 142 so far does not contain interference with the Standard Model $\mathrm{W}^{\pm}$ --- in practice this should not be a major limitation. The couplings of the $\mathrm{W}'$ to quarks and leptons are set via \ttt{PARU(131) - PARU(134)}. Again one may set vector and axial couplings freely, separately for the $\mathrm{q} \overline{\mathrm{q}}'$ and the $\ell \nu_{\ell}$ decay channels. The defaults correspond to the $V-A$ structure of the Standard Model $\mathrm{W}$, but can be changed to simulate a wide selection of models. One possible limitation is that the same Cabibbo--Kobayashi--Maskawa quark mixing matrix is assumed as for the standard $\mathrm{W}$. The coupling $\mathrm{W}' \to \mathrm{Z}^0 \mathrm{W}$ can be set via \ttt{PARU(135) - PARU(136)}. Further comments on this channel as for $\mathrm{Z}'$; in particular, default couplings again agree with the `extended gauge model' of \cite{Alt89}. A $\mathrm{W}' \to \mathrm{W} \H^0$ channel is also included, in analogy with the $\mathrm{Z}'^0 \to \mathrm{Z}^0 \H^0$ one, see \ttt{PARU(146)}. The $\mathrm{R}$ boson (particle code 40) of subprocess 144 represents one possible scenario for a horizontal gauge boson, i.e. a gauge boson that couples between the generations, inducing processes like $\mathrm{s} \overline{\mathrm{d}} \to \mathrm{R}^0 \to \mu^- \mathrm{e}^+$. Experimental limits on flavour-changing neutral currents forces such a boson to be fairly heavy. The model implemented is the one described in \cite{Ben85a}. \subsubsection{Leptoquarks} \label{sss:LQclass} \ttt{MSEL} = 25 \\ ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 145 & $\mathrm{q}_i \ell_j \to \L_{\mathrm{Q}}$ \\ 162 & $\mathrm{q} \mathrm{g} \to \ell \L_{\mathrm{Q}}$ \\ 163 & $\mathrm{g} \mathrm{g} \to \L_{\mathrm{Q}} \br{\L}_{\mathrm{Q}}$ \\ 164 & $\mathrm{q}_i \overline{\mathrm{q}}_i \to \L_{\mathrm{Q}} \br{\L}_{\mathrm{Q}}$ \\ \end{tabular} Several processes that can generate a leptoquark have been included. Currently only one leptoquark has been implemented, as particle 39, denoted $\L_{\mathrm{Q}}$. The leptoquark is assumed to carry specific quark and lepton quantum numbers, by default $\u$ quark plus electron. These flavour numbers are conserved, i.e. a process such as $\u \mathrm{e}^- \to \L_{\mathrm{Q}} \to \d \nu_{\mathrm{e}}$ is not allowed. This may be a bit restrictive, but it represents one of many leptoquark possibilities. The spin of the leptoquark is assumed to be zero, i.e. its decay is isotropical. Although only one leptoquark is implemented, its flavours may be changed arbitrarily to study the different possibilities. The flavours of the leptoquark are defined by the quark and lepton flavours in the decay mode list. Since only one decay channel is allowed, this means that the quark flavour is stored in \ttt{KFDP(MDCY(39,2),1)} and the lepton one in \ttt{KFDP(MDCY(39,2),2)}. The former must always be a quark, while the latter could be a lepton or an antilepton; a charge-conjugate partner is automatically defined by the program. At initialization, the charge is recalculated as a function of the flavours defined; also the leptoquark name is redefined to be of the type \ttt{'LQ\_(q)(l)'}, where actual quark \ttt{(q)} and lepton \ttt{(l)} flavours are displayed. The $\L_{\mathrm{Q}} \to \mathrm{q} \ell$ vertex contains an undetermined Yukawa coupling strength, which affects both the width of the leptoquark and the cross section for many of the production graphs. This strength may be changed in \ttt{PARU(151)}. The definition of \ttt{PARU(151)} corresponds to the $k$ factor of \cite{Hew88}, i.e. to $\lambda^2/(4\pi\alpha_{\mathrm{em}})$, where $\lambda$ is the Yukawa coupling strength of \cite{Wud86}. Note that \ttt{PARU(151)} is thus quadratic in the coupling. The leptoquark is likely to be fairly long-lived, in which case it has time to fragment into a mesonic- or baryonic-type state, which would decay later on. This is a bit tedious to handle; therefore the leptoquark is always assumed to decay before fragmentation has to be considered. This may give some imperfections in the event generation, but should not be off by much in the final analysis. Inside the program, the leptoquark is treated as a resonance. Since it carries colour, some extra care is required. In particular, it is not allowed to put the leptoquark stable, by modifying either \ttt{MDCY(39,1)} or \ttt{MSTP(41)}: then the leptoquark would be handed undecayed to {\tsc{Jetset}}, which would try to fragment it (as it does with any other coloured object), and most likely crash. \subsubsection{Compositeness and anomalous couplings} \label{sss:ancoupclass} ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 11 & $\mathrm{f}_i \mathrm{f}_j \to \mathrm{f}_i \mathrm{f}_j$ (QCD) \\ 12 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{f}_k \overline{\mathrm{f}}_k$ \\ 20 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \gamma \mathrm{W}^+$ \\ 165 & $\mathrm{f}_i \overline{\mathrm{f}}_i \to \mathrm{f}_k \overline{\mathrm{f}}_k$ (via $\gamma^* / \Z^0$) \\ 166 & $\mathrm{f}_i \overline{\mathrm{f}}_j \to \mathrm{f}_k \overline{\mathrm{f}}_l$ (via $\mathrm{W}^{\pm}$) \\ \end{tabular} Some processes have been set up to allow anomalous coupling to be introduced, in addition to the Standard Model ones. These can be switched on by \ttt{MSTP(5)}$\geq 1$; the default \ttt{MSTP(5)=0} corresponds to the Standard Model behaviour. In processes 11 and 12, the quark substructure is included in the left--left isoscalar model \cite{Eic84,Chi90} for \ttt{MSTP(5)=1}, with compositeness scale $\Lambda$ given in \ttt{PARU(155)} (default 1000~GeV) and sign $\eta$ of interference term in \ttt{PARU(156)} (default $+1$; only other alternative $-1$). The above model assumes that only $\u$ and $\d$ quarks are composite (at least at the scale studied); with \ttt{MSTP(5)=2} compositeness terms are included in the interactions between all quarks. The processes 165 and 166 are basically equivalent to 1 and 2, i.e. $\gamma^* / \Z^0$ and $\mathrm{W}^{\pm}$ exchange, respectively, but a bit less fancy (no mass-dependent width etc.). The reason for this duplication is that the resonance treatment formalism of processes 1 and 2 could not easily be extended to include other than $s$-channel graphs. In processes 165 and 166, only one final-state flavour is generated at the time; this flavour should be set in \ttt{KFPR(165,1)} and \ttt{KFPR(166,1)}, respectively. For process 166 one gives the down-type flavour, and the program will associate the up-type flavour of the same generation. Defaults are 11 in both cases, i.e. $\e^+\e^-$ and $\mathrm{e}^+ \nu_{\mathrm{e}}$ ($\mathrm{e}^- \br{\nu}_{\mathrm{e}}$) final states. While \ttt{MSTP(5)=0} gives the Standard Model results, \ttt{MSTP(5)=1} contains the left--left isoscalar model (which does not affect process 166), and \ttt{MSTP(5)=3} the helicity-non-conserving model (which affects both) \cite{Eic84,Lan91}. Both models above assume that only $\u$ and $\d$ quarks are composite; with \ttt{MSTP(5)=} 2 or 4, respectively, contact terms are included for all quarks in the initial state. Parameters are \ttt{PARU(155)} and \ttt{PARU(156)}, as above. Note that processes 165 and 166 are bookkept as $2 \to 2$ processes, while 1 and 2 are $2 \to 1$ ones. This means that the default $\mathrm{Q}^2$ scale in parton distributions is $p_{\perp}^2$ for the former and $\hat{s}$ for the latter. To make contact between the two, it is recommended to set \ttt{MSTP(32)=4}, so as to use $\hat{s}$ as scale also for processes 165 and 166. In process 20, for $\mathrm{W} \gamma$ pair production, it is possible to set an anomalous magnetic moment for the $\mathrm{W}$ in \ttt{PARU(153)} ($= \eta = \kappa-1$; where $\kappa = 1$ is the Standard Model value). The production process is affected according to the formulae of \cite{Sam91}, while $\mathrm{W}$ decay currently remains unaffected. It is necessary to set \ttt{MSTP(5)=1} to enable this extension. \subsubsection{Excited fermions} \label{sss:qlstarclass} ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 147 & $\d \mathrm{g} \to \d^*$ \\ 148 & $\u \mathrm{g} \to \u^*$ \\ 167 & $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \d^*$ \\ 168 & $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \u^*$ \\ \end{tabular} Compositeness scenarios may also give rise to sharp resonances of excited quarks and leptons. If \ttt{MSTP(6)=1}, then at initialization the standard fourth generation of fermions will be overwritten, and made to correspond to an excited copy of the first generation, consisting of spin $1/2$ particles $\d^*$ (code 7), $\u^*$ (8), $\mathrm{e}^*$ (17) and $\nu^*_{\mathrm{e}}$ (18). Since the original fourth-generation information is lost, it is then not possible to generate fourth-generation particles in the same run. The current implementation contains gauge interaction production by quark--gluon fusion (processes 147 and 148) and contact interaction production by quark--quark or quark--antiquark scattering (processes 167 and 168) . The couplings $f$, $f'$ and $f_s$ to the {\bf SU(2)}, {\bf U(1)} and {\bf SU(3)} groups are stored in \ttt{PARU(157) - PARU(159)}, the scale parameter $\Lambda$ in \ttt{PARU(155)}; you are also expected to change the $\mathrm{f}^*$ masses in accordance with what is desired --- see \cite{Bau90} for details on conventions. Decay processes are of the types $\mathrm{q}^* \to \mathrm{q} \mathrm{g}$, $\mathrm{q}^* \to \mathrm{q} \gamma$, $\mathrm{q}^* \to \mathrm{q} \mathrm{Z}^0$ or $\mathrm{q}^* \to \mathrm{q}' \mathrm{W}^{\pm}$. A non-trivial angular dependence is included in the $\mathrm{q}^*$ decay for processes 147 and 148, but has not been included for processes 167 and 168. \subsubsection{Technicolor} \label{sss:technicolorclass} ISUB = \\ \begin{tabular}{rl@{\protect\rule{0mm}{\tablinsep}}} 149 & $\mathrm{g} \mathrm{g} \to \eta_{\mrm{techni}}$ \\ \end{tabular} The technicolor scenario offers an alternative to the ordinary Higgs mechanism for giving masses to the $\mathrm{W}$ and $\mathrm{Z}$. The technicolor gauge group is an analogue of QCD, with a rich spectrum of technimesons made out of techniquarks. Three of the technipions assume the role of the longitudinal components of the $\mathrm{W}$ and $\mathrm{Z}$ bosons, but many other states remain as separate particles. No fully realistic model has been found so far, however, so any phenomenology has to be taken as indicative only. In section \ref{sss:heavySMHclass} it is discussed how processes 71--77, in some of its options, can be used to simulate a scenario with techni-$\rho$ resonances in longitudinal gauge boson scattering. Here we present another process, that of the production of a techni-$\eta$. This particle has zero spin, is a singlet under electroweak {\bf SU(2)}$\times${\bf U(1)}, but carries octet colour charge. It is one of the possible techni-$\pi$ particles; the name techni-$\eta$ is part of a subclassification not used by all authors. The techni-$\eta$ couples to ordinary fermions according to the fermion squared mass. The dominant decay mode is therefore $\t \overline{\mathrm{t}}$, if allowed. The coupling to a $\mathrm{g} \mathrm{g}$ state is roughly comparable with that to $\b \overline{\mathrm{b}}$. Production at hadron colliders is therefore predominantly through $\mathrm{g} \mathrm{g}$ fusion, as implemented in process 149. The two main free parameters are the techni-$\eta$ mass and the decay constant $F_{\pi}$. The latter appears inversely quadratically in all the partial widths. Also the total cross section is affected, since the cross section is proportional to the $\mathrm{g} \mathrm{g}$ partial width. $F_{\pi}$ is stored in \ttt{PARP(46)} and has the default value 123 GeV, which is the number predicted in some models. \subsection{Main Processes by Machine} In the previous section we have already commented on which processes have limited validity, or have different meanings (according to conventional terminology) in different contexts. Let us just repeat a few of the main points to be remembered for different machines. \subsubsection{$\e^+\e^-$ collisions} The main annihilation process is number 1, $\e^+\e^- \to \mathrm{Z}^0$, where in fact the full $\gamma^*/\mathrm{Z}^0$ interference structure is included. This process can be used, with some confidence, for c.m. energies from about 4 GeV upwards, i.e. at DORIS/CESR, PETRA/PEP, TRISTAN, LEP, and any future linear colliders. (To get below 10 GeV, you have to change \ttt{PARP(2)}, however.) This is the default process obtained when \ttt{MSEL=1}, i.e. when you do not change anything yourself. Process 141 contains a $\mathrm{Z}'^0$, including full interference with the standard $\gamma^* / \Z^0$. With the value \ttt{MSTP(44)=4} in fact one is back at the standard $\gamma^* / \Z^0$ structure, i.e. the $\mathrm{Z}'^0$ piece has been switched off. Even so, this process may be useful, since it can simulate e.g. $\e^+\e^- \to \H^0 \mathrm{A}^0$. Since the $\H^0$ may in its turn decay to $\mathrm{Z}^0 \mathrm{Z}^0$, a decay channel of the ordinary $\mathrm{Z}^0$ to $\H^0 \mathrm{A}^0$, although physically correct, would be technically confusing. In particular, it would be messy to set the original $\mathrm{Z}^0$ to decay one way and the subsequent ones another. So, in this sense, the $\mathrm{Z}'^0$ could be used as a copy of the ordinary $\mathrm{Z}^0$, but with a distinguishable label. The process $\e^+\e^- \to \Upsilon$ does not exist as a separate process in {\tsc{Pythia}}, but can be simulated by using \ttt{LUONIA}, see section \ref{ss:oniadecays}. At LEP 2 and even higher energy machines, the simple $s$-channel process 1 will lose out to other processes, such as $\e^+\e^- \to \mathrm{Z}^0 \mathrm{Z}^0$ and $\e^+\e^- \to \mathrm{W}^+ \mathrm{W}^-$, i.e. processes 22 and 25. The former process in fact includes the structure $\e^+\e^- \to (\gamma^* / \Z^0)(\gamma^* / \Z^0)$, which means that the cross section is singular if either of the two $\gamma^* / \Z^0$ masses is allowed to vanish. A mass cut therefore needs to be introduced, and is actually also used in other processes, such as $\e^+\e^- \to \mathrm{W}^+ \mathrm{W}^-$. For practical applications, both with respect to cross sections and to event shapes, it is imperative to include initial-state radiation effects. Therefore \ttt{MSTP(11)=1} is the default, wherein exponentiated electron-inside-electron distributions are used to give the momentum of the actually interacting electron. By radiative corrections to process 1, such processes as $\e^+\e^- \to \gamma \mathrm{Z}^0$ are therefore automatically generated. If process 19 were to be used at the same time, this would mean that radiation were to be double-counted. In the alternative \ttt{MSTP(11)=0}, electrons are assumed to deposit their full energy in the hard process, i.e. initial-state QED radiation is not included. This option is very useful, since it often corresponds to the `ideal' events that one wants to correct back to. Resolved electrons also means that one may have interactions between photons. This opens up the whole field of $\gamma\gamma$ processes, which is described in section \ref{sss:photoprodclass}. In particular, with \ttt{MSTP(12)=1} photons may be resolved, i.e. photons need not only interact point-like, but can also interact like a hadron with a partonic substructure. The whole menagerie of hadron--hadron collider processes can then be accessed. However, it is not yet possible to include the low-$p_{\perp}$ processes with a variable photon energy spectrum. That is, to generate the `total' $\gamma\gamma$ spectrum, the program also has to be initialized for a $\gamma\gamma$ collider. The thrust of the {\tsc{Pythia/Jetset}} programs is towards processes that involve hadron production, one way or another. Because of generalizations from other areas, also a few completely non-hadronic processes are available. These include Bhabha scattering, $\e^+\e^- \to \e^+\e^-$ in process 10, and photon pair production, $\e^+\e^- \to \gamma \gamma$ in process 18. However, note that the precision that could be expected in a {\tsc{Pythia}} simulation of those processes is certainly far less than that of dedicated programs. For one thing, electroweak loop effects are not included. For another, nowhere is the electron mass taken into account, which means that explicit cut-offs at some minimum $p_{\perp}$ are always necessary. \subsubsection{Lepton--hadron collisions} The issue of applications to $\mathrm{e}\mathrm{p}$ colliders has been covered in a recent report \cite{Sjo92b}. The default process for a lepton--hadron collider is deep inelastic scattering, $\ell \mathrm{q} \to \ell' \mathrm{q}'$, of process 10. This includes $\gamma^0/\mathrm{Z}^0/\mathrm{W}^{\pm}$ exchange, with full interference, as described in section \ref{sss:DISclass}. Radiation off the incoming lepton leg is included by \ttt{MSTP(11)=1} and off the outgoing one by \ttt{MSTJ(41)=2} (both are default). Note that both QED and QCD radiation (off the $\mathrm{e}$ and the $\mathrm{q}$ legs, respectively) are allowed to modify the $x$ and $Q^2$ values of the process, while the conventional approach in the literature is to allow only the former. Therefore an option (on by default) has been added to preserve these values by a post-facto rescaling, \ttt{MSTP(23)=1}. In terms of cross sections, a more important set of processes are related to photoproduction, either with a point-like or with a resolved photon, see section \ref{sss:photoprodclass}. A complete description of photoproduction is available \cite{Sch93,Sch93a}, but needs three separate runs for the three distinct behaviours of a photon: point-like, VMD resolved and anomalous resolved. \subsubsection{Hadron--hadron collisions} The default is to include QCD jet production by $2 \to 2$ processes, see section \ref{sss:QCDjetclass}. Since the differential cross section is divergent for $p_{\perp} \to 0$, a lower cut-off has to be introduced. Normally that cut-off is given by the user-set $p_{\perp\mathrm{min}}$ value in \ttt{CKIN(3)}. If \ttt{CKIN(3)} is chosen smaller than a given value of the order of 2 GeV (see \ttt{PARP(81)} and \ttt{PARP(82)}), then low-$p_{\perp}$ events are also switched on. The jet cross section is regularized at low $p_{\perp}$, so as to obtain a smooth joining between the high-$p_{\perp}$ and the low-$p_{\perp}$ descriptions, see further section \ref{ss:multint}. As \ttt{CKIN(3)} is varied, the jump from one scenario to another is abrupt, in terms of cross section: in a high-energy hadron collider, the cross section for jets down to a $p_{\perp\mathrm{min}}$ scale of a few GeV can well reach values much larger than the total inelastic, non-diffractive cross section. Clearly this is nonsense; therefore either $p_{\perp\mathrm{min}}$ should be picked so large that the jet cross section be only a fraction of the total one, or else one should select $p_{\perp\mathrm{min}} = 0$ and make use of the full description. If one switches to \ttt{MSEL=2}, also elastic and diffractive processes are switched on, see section \ref{sss:minbiasclass}. However, the simulation of these processes is fairly primitive, and should not be used for dedicated studies, but only to estimate how much they may contaminate the class of non-diffractive minimum bias events. Most processes can be simulated in hadron colliders, since the bulk of {\tsc{Pythia}} processes can be initiated by quarks or gluons. However, there are limits. Currently we include no photon or lepton parton distributions, which means that a process like $\gamma \mathrm{q} \to \gamma \mathrm{q}$ is not accessible. Further, the possibility of having $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$ interacting in processes such as 71--77 has been hardwired process by process, and does not mean that there is a generic treatment of $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$ distributions. The emphasis in the hadron--hadron process description is on high energy hadron colliders. The program can be used also at fixed-target energies, but the multiple interaction model for underlying events then breaks down and should not be used. The limit of applicability is somewhere at around 100 GeV. Below that, one is also recommended to change to \ttt{MSTP(92)=3}, to obtain a reasonable amount of beam remnant particle production in the absence of multiple interactions. \clearpage \section{The PYTHIA Program Elements} In the previous two sections, the physics processes and the event-generation schemes of {\tsc{Pythia}} have been presented. Here, finally, the event-generation routines and the common block variables are described. However, routines and variables related to initial- and final-state showers, beam remnants and underlying events, and fragmentation and decay are relegated to subsequent sections on these topics. Further, for historical reasons, many adjustable coupling constants are found in the \ttt{LUDAT1} common block in {\tsc{Jetset}}, rather than somewhere in the {\tsc{Pythia}} common blocks; these parameters are described in section \ref{ss:coupcons}. In the presentation in this section, information less important for an efficient use of {\tsc{Pythia}} has been put closer to the end. We therefore begin with the main event generation routines, and follow this by the main common block variables. It is useful to distinguish three phases in a normal run with {\tsc{Pythia}}. In the first phase, the initialization, the general character of the run is determined. At a minimum, this requires the specification of the incoming hadrons and the energies involved. At the discretion of the user, it is also possible to select specific final states, and to make a number of decisions about details in the subsequent generation. This step is finished by a \ttt{PYINIT} call, at which time several variables are initialized in accordance with the values set. The second phase consists of the main loop over the number of events, with each new event being generated by a \ttt{PYEVNT} call. This event may then be analysed, using information stored in some common blocks, and the statistics accumulated. In the final phase, results are presented. This may often be done without the invocation of any {\tsc{Pythia}} routines. From \ttt{PYSTAT}, however, it is possible to obtain a useful list of cross sections for the different subprocesses. \subsection{The Main Subroutines} \label{ss:PYTmainroutines} There are two routines that you must know: \ttt{PYINIT} for initialization and \ttt{PYEVNT} for the subsequent generation of each new event. In addition, the cross section and other kinds of information available with \ttt{PYSTAT} are frequently useful. The other three routines described here, \ttt{PYFRAM}, \ttt{PYKCUT}, and \ttt{PYEVWT}, are of more specialized interest. \drawbox{CALL PYINIT(FRAME,BEAM,TARGET,WIN)}\label{p:PYINIT} \begin{entry} \itemc{Purpose:} to initialize the generation procedure. \iteme{FRAME :} a character variable used to specify the frame of the experiment. Upper-case and lower-case letters may be freely mixed. \begin{subentry} \iteme{= 'CMS' :} colliding beam experiment in c.m. frame, with beam momentum in $+z$ direction and target momentum in $-z$ direction. \iteme{= 'FIXT' :} fixed-target experiment, with beam particle momentum pointing in $+z$ direction. \iteme{= 'USER' :} full freedom to specify frame by giving beam momentum in \ttt{P(1,1)}, \ttt{P(1,2)} and \ttt{P(1,3)} and target momentum in \ttt{P(2,1)}, \ttt{P(2,2)} and \ttt{P(2,3)} in common block \ttt{LUJETS}. Particles are assumed on the mass shell, and energies are calculated accordingly. \iteme{= 'FOUR' :} as \ttt{'USER'}, except also energies should be specified, in \ttt{P(1,4)} and \ttt{P(2,4)}, respectively. The particles need not be on the mass shell; effective masses are calculated from energy and momentum. (But note that numerical precision may suffer; if you know the masses the option \ttt{'FIVE'} below is preferrable.) \iteme{= 'FIVE' :} as \ttt{'USER'}, except also energies and masses should be specified, i.e the full momentum information in \ttt{P(1,1) - P(1,5)} and \ttt{P(2,1) - P(2,5)} should be given for beam and target, respectively. Particles need not be on the mass shell. Space-like virtualities should be stored as $-\sqrt{-m^2}$. Four-momentum and mass information must match. \iteme{= 'NONE' :} there will be no initialization of any processes, but only of resonance widths and a few other process-independent variables. Subsequent to such a call, \ttt{PYEVNT} cannot be used to generate events, so this option is mainly intended for those who will want to construct their own events afterwards, but still want to have access to some of the {\tsc{Pythia}} facilities. In this option, the \ttt{BEAM}, \ttt{TARGET} and \ttt{WIN} arguments are dummy. \end{subentry} \iteme{BEAM, TARGET :} character variables to specify beam and target particles. Upper-case and lower-case letters may be freely mixed. An antiparticle may be denoted either by `$\sim$' or `bar' at the end of the name. It is also possible to leave out the underscore (`\_') directly after `nu' in neutrino names, and the charge for proton and neutron. \begin{subentry} \iteme{= 'e-' :} electron. \iteme{= 'e+' :} positron. \iteme{= 'nu\_e' :} $\nu_{\mathrm{e}}$. \iteme{= 'nu\_e$\sim$' :} $\br{\nu}_{\mathrm{e}}$. \iteme{= 'mu-' :} $\mu^-$. \iteme{= 'mu+' :} $\mu^+$. \iteme{= 'nu\_mu' :} $\nu_{\mu}$. \iteme{= 'nu\_mu$\sim$' :} $\br{\nu}_{\mu}$. \iteme{= 'tau-' :} $\tau^-$. \iteme{= 'tau+' :} $\tau^+$. \iteme{= 'nu\_tau' :} $\nu_{\tau}$. \iteme{= 'nu\_tau$\sim$' :} $\br{\nu}_{\tau}$. \iteme{= 'gamma' :} photon (real, i.e. on the mass shell). \iteme{= 'pi0' :} $\pi^0$. \iteme{= 'pi+' :} $\pi^+$. \iteme{= 'pi-' :} $\pi^-$. \iteme{= 'n0' :} neutron. \iteme{= 'n$\sim$0' :} antineutron. \iteme{= 'p+' :} proton. \iteme{= 'p$\sim$-' :} antiproton. \iteme{= 'Lambda0' :} $\Lambda$ baryon. \iteme{= 'Sigma-' :} $\Sigma^-$ baryon. \iteme{= 'Sigma0' :} $\Sigma^0$ baryon. \iteme{= 'Sigma+' :} $\Sigma^+$ baryon. \iteme{= 'Xi-' :} $\Xi^-$ baryon. \iteme{= 'Xi0' :} $\Xi^0$ baryon. \iteme{= 'Omega-' :} $\Omega^-$ baryon. \iteme{= 'pomeron' :} the pomeron $\mathrm{I}\!\mathrm{P}$; since pomeron parton distribution functions have not been defined this option can not be used currently. \iteme{= 'reggeon' :} the reggeon $\mathrm{I}\!\mathrm{R}$, with comments as for the pomeron above. \end{subentry} \iteme{WIN :} related to energy of system, exact meaning depends on \ttt{FRAME}. \begin{subentry} \iteme{FRAME='CMS' :} total energy of system (in GeV). \iteme{FRAME='FIXT' :} momentum of beam particle (in GeV/$c$). \iteme{FRAME='USER' :} dummy (information is taken from the \ttt{P} vectors, see above). \end{subentry} \end{entry} \drawbox{CALL PYEVNT}\label{p:PYEVNT} \begin{entry} \itemc{Purpose:} to generate one event of the type specified by the \ttt{PYINIT} call. (This is the main routine, which calls a number of other routines for specific tasks.) \end{entry} \drawbox{CALL PYSTAT(MSTAT)}\label{p:PYSTAT} \begin{entry} \itemc{Purpose:} to print out cross-sections statistics, decay widths, branching ratios, status codes and parameter values. \ttt{PYSTAT} may be called at any time, after the \ttt{PYINIT} call, e.g. at the end of the run, or not at all. \iteme{MSTAT :} specification of desired information. \begin{subentry} \iteme{= 1 :} prints a table of how many events of the different kinds that have been generated and the corresponding cross sections. All numbers already include the effects of cuts required by the user in \ttt{PYKCUT}. \iteme{= 2 :} prints a table of the resonances defined in the program, with their particle codes (KF), and all allowed decay channels. (If the number of generations in \ttt{MSTP(1)} is 3, however, channels involving fourth-generation particles are not displayed.) For each decay channel is shown the sequential channel number (IDC) of the {\tsc{Jetset}} decay tables, the partial decay width, branching ratio and effective branching ratio (in the event some channels have been excluded by the user). \iteme{= 3 :} prints a table with the allowed hard interaction flavours \ttt{KFIN(I,J)} for beam and target particles. \iteme{= 4 :} prints a table of the kinematical cuts \ttt{CKIN(I)} set by the user in the current run. \iteme{= 5 :} prints a table with all the values of the status codes \ttt{MSTP(I)} and the parameters \ttt{PARP(I)} used in the current run. \end{subentry} \end{entry} \drawbox{CALL PYFRAM(IFRAME)}\label{p:PYFRAM} \begin{entry} \itemc{Purpose:} to transform an event listing between different reference frames, if so desired. \iteme{IFRAME :} specification of frame the event is to be boosted to. \begin{subentry} \iteme{= 1 :} frame specified by user in the \ttt{PYINIT} call. \iteme{= 2 :} c.m. frame of incoming particles. \iteme{= 3 :} hadronic c.m. frame of lepton--hadron interaction events. Mainly intended for deep inelastic scattering, but can also be used in photoproduction. Note that both the lepton and any photons radiated off the lepton remain in the event listing, and have to be removed separately if you only want to study the hadronic subsystem. \end{subentry} \end{entry} \drawbox{CALL PYKCUT(MCUT)}\label{p:PYKCUT} \begin{entry} \itemc{Purpose:} to enable you to reject a given set of kinematic variables at an early stage of the generation procedure (before evaluation of cross sections), so as not to spend unnecessary time on the generation of events that are not wanted. The routine will not be called unless you require is by setting \ttt{MSTP(141)=1}, and never if `minimum-bias'-type events (including elastic and diffractive scattering) are to be generated as well. A dummy routine \ttt{PYKCUT} is included in the program file, so as to avoid unresolved external references when the routine is not used. \iteme{MCUT :} flag to signal effect of user-defined cuts. \begin{subentry} \iteme{= 0 :} event is to be retained and generated in full. \iteme{= 1 :} event is to be rejected and a new one generated. \end{subentry} \itemc{Remark :} at the time of selection, several variables in the \ttt{MINT} and \ttt{VINT} arrays in the \ttt{PYINT1} common block contain information that can be used to make the decision. The routine provided in the program file explicitly reads the variables that have been defined at the time \ttt{PYKCUT} is called, and also calculates some derived quantities. The information available includes subprocess type ISUB, $E_{\mrm{cm}}$, $\hat{s}$, $\hat{t}$, $\hat{u}$, $\hat{p}_{\perp}$, $x_1$, $x_2$, $x_{\mrm{F}}$, $\tau$, $y$, $\tau'$, $\cos\hat{\theta}$, and a few more. Some of these may not be relevant for the process under study, and are then set to zero. \end{entry} \drawbox{CALL PYEVWT(WTXS)}\label{p:PYEVWT} \begin{entry} \itemc{Purpose:} to allow you to reweight event cross sections, by process type and kinematics of the hard scattering. There exists two separate modes of usage, described in the following. \\ For \ttt{MSTP(142)=1}, it is assumed that the cross section of the process is correctly given by default in {\tsc{Pythia}}, but that one wishes to generate events biased to a specific region of phase space. While the \ttt{WTXS} factor therefore multiplies the na\"{\i}ve cross section in the choice of subprocess type and kinematics, the produced event comes with a compensating weight \ttt{PARI(10)=1./WTXS}, which should be used when filling histograms etc. In the \ttt{PYSTAT(1)} table, the cross sections are unchanged (up to statistical errors) compared with the standard cross sections, but the relative composition of events may be changed and need no longer be in proportion to relative cross sections. A typical example of this usage is if one wishes to enhance the production of high-$p_{\perp}$ events; then a weight like \ttt{WTXS}$=(p_{\perp}/p_{\perp 0})^2$ (with $p_{\perp 0}$ some fixed number) might be appropriate. \\ For \ttt{MSTP(142)=2}, on the other hand, it is assumed that the true cross section is really to be modifed by the multiplicative factor \ttt{WTXS}. The generated events therefore come with unit weight, just as usual. This option is really equivalent to replacing the basic cross sections coded in {\tsc{Pythia}}, but allows more flexibility: no need to recompile the whole of {\tsc{Pythia}}. \\ The routine will not be called unless \ttt{MSTP(142)}$\geq 1$, and never if `minimum-bias'-type events (including elastic and diffractive scattering) are to be generated as well. Further, cross sections for additional multiple interactions or pile-up events are never affected. A dummy routine \ttt{PYEVWT} is included in the program file, so as to avoid unresolved external references when the routine is not used. \iteme{WTXS:} multiplication factor to ordinary event cross section; to be set (by you) in \ttt{PYEVWT} call. \itemc{Remark :} at the time of selection, several variables in the \ttt{MINT} and \ttt{VINT} arrays in the \ttt{PYINT1} common block contain information that can be used to make the decision. The routine provided in the program file explicitly reads the variables that have been defined at the time \ttt{PYEVWT} is called, and also calculates some derived quantities. The given list of information includes subprocess type ISUB, $E_{\mrm{cm}}$, $\hat{s}$, $\hat{t}$, $\hat{u}$, $\hat{p}_{\perp}$, $x_1$, $x_2$, $x_{\mrm{F}}$, $\tau$, $y$, $\tau'$, $\cos\hat{\theta}$, and a few more. Some of these may not be relevant for the process under study, and are then set to zero. \itemc{Warning:} the weights only apply to the hard scattering subprocesses. There is no way to reweight the shape of initial- and final-state showers, fragmentation, or other aspects of the event. \end{entry} \subsection{Switches for Event Type and Kinematics Selection} \label{ss:PYswitchkin} By default, if {\tsc{Pythia}} is run for a hadron collider, only QCD $2 \to 2$ processes are generated, composed of hard interactions above $p_{\perp\mathrm{min}}=$\ttt{PARP(81)}, with low-$p_{\perp}$ processes added on so as to give the full (parametrized) inelastic, non-diffractive cross section. In an $\e^+\e^-$ collider, $\gamma^* / \Z^0$ production is the default, and in an $\e\p$ one it is deep inelastic scattering. With the help of the common block \ttt{PYSUBS}, it is possible to select the generation of another process, or combination of processes. It is also allowed to restrict the generation to specific incoming partons/particles at the hard interaction. This often automatically also restricts final-state flavours but, in processes such as resonance production or QCD/QED production of new flavours, switches in the {\tsc{Jetset}} program may be used to this end; see section \ref{ss:parapartdat}. The \ttt{CKIN} array may be used to impose specific kinematics cuts. You should here be warned that, if kinematical variables are too strongly restricted, the generation time per event may become very long. In extreme cases, where the cuts effectively close the full phase space, the event generation may run into an infinite loop. The generation of $2 \to 1$ resonance production is performed in terms of the $\hat{s}$ and $y$ variables, and so the ranges \ttt{CKIN(1) - CKIN(2)} and \ttt{CKIN(7) - CKIN(8)} may be arbitrarily restricted without a significant loss of speed. For $2 \to 2$ processes, $\cos\hat{\theta}$ is added as a third generation variable, and so additionally the range \ttt{CKIN(27) - CKIN(28)} may be restricted without any danger. Effects from initial- and final-state radiation are not included, since they are not known at the time the kinematics at the hard interaction is selected. The sharp kinematical cut-offs that can be imposed on the generation process are therefore smeared, both by QCD radiation and by fragmentation. A few examples of such effects follow. \begin{Itemize} \item Initial-state radiation implies that each of the two incoming partons has a non-vanishing $p_{\perp}$ when they interact. The hard scattering subsystem thus receives a net transverse boost, and is rotated with respect to the beam directions. In a $2 \to 2$ process, what typically happens is that one of the scattered partons receives an increased $p_{\perp}$, while the $p_{\perp}$ of the other parton is reduced. \item Since the initial-state radiation machinery assigns space-like virtualities to the incoming partons, the definitions of $x$ in terms of energy fractions and in terms of momentum fractions no longer coincide, and so the interacting subsystem may receive a net longitudinal boost compared with na\"{\i}ve expectations, as part of the parton-shower machinery. \item Initial-state radiation gives rise to additional jets, which in extreme cases may be mistaken for either of the jets of the hard interaction. \item Final-state radiation gives rise to additional jets, which smears the meaning of the basic $2 \to 2$ scattering. The assignment of soft jets is not unique. The energy of a jet becomes dependent on the way it is identified, e.g. what jet cone size is used. \item The beam remnant description assigns primordial $k_{\perp}$ values, which also gives a net $p_{\perp}$ shift of the hard-interaction subsystem; except at low energies this effect is overshadowed by initial-state radiation, however. Beam remnants may also add further activity under the `perturbative' event. \item Fragmentation will further broaden jet profiles, and make jet assignments and energy determinations even more uncertain. \end{Itemize} In a study of events within a given window of experimentally defined variables, it is up to you to leave such liberal margins that no events are missed. In other words, cuts have to be chosen such that a negligible fraction of events migrate from outside the simulated region to inside the interesting region. Often this may lead to low efficiency in terms of what fraction of the generated events are actually of interest to you. See also section \ref{ss:PYTstarted}. In addition to the variables found in \ttt{PYSUBS}, also those in the \ttt{PYPARS} common block may be used to select exactly what one wants to have simulated. These possibilities will be described in the following subsection. The notation used above and in the following is that `$\hat{~}$' denotes internal variables in the hard scattering subsystem, while `$^*$' is for variables in the c.m. frame of the event as a whole. \drawbox{COMMON/PYSUBS/MSEL,MSUB(200),KFIN(2,-40:40),CKIN(200)}% \label{p:PYSUBS} \begin{entry} \itemc{Purpose:} to allow the user to run the program with any desired subset of processes, or restrict flavours or kinematics. If the default values, denoted below by (D=\ldots), are not satisfactory, they must be changed before the \ttt{PYINIT} call. \boxsep \iteme{MSEL :}\label{p:MSEL} (D=1) a switch to select between full user control and some preprogrammed alternatives. \begin{subentry} \iteme{= 0 :} desired subprocesses have to be switched on in \ttt{MSUB}, i.e. full user control. \iteme{= 1 :} depending on incoming particles, different alternatives are used. \\ Lepton--lepton: $\mathrm{Z}$ or $\mathrm{W}$ production (ISUB = 1 or 2). \\ Lepton--hadron: deep inelastic scattering (ISUB = 10). \\ Hadron--hadron: QCD high-$p_{\perp}$ processes (ISUB = 11, 12, 13, 28, 53, 68); additionally low-$p_{\perp}$ production if \ttt{CKIN(3)}$<$\ttt{PARP(81)} or \ttt{PARP(82)}, depending on \ttt{MSTP(82)} (ISUB = 95). If low-$p_{\perp}$ is switched on, the other \ttt{CKIN} cuts are not used. \\ A resolved photon counts as hadron, except that an anomalous photon cannot have low-$p_{\perp}$ interactions. When the photon is not resolved, the following cases are possible. \\ Photon--lepton: Compton scattering (ISUB = 34). \\ Photon--hadron: photon-parton scattering (ISUB = 33, 34, 54). \\ Photon--photon: fermion pair production (ISUB = 58). \iteme{= 2 :} as \ttt{MSEL = 1} for lepton--lepton, lepton--hadron and unresolved photons. For hadron--hadron (including resolved photons) all QCD processes, including low-$p_{\perp}$, single and double diffractive and elastic scattering, are included (ISUB = 11, 12, 13, 28, 53, 68, 91, 92, 93, 94, 95). The \ttt{CKIN} cuts are here not used. \iteme{= 4 :} charm ($\c\overline{\mathrm{c}}$) production with massive matrix elements (ISUB = 81, 82, 84, 85). \iteme{= 5 :} bottom ($\b\overline{\mathrm{b}}$) production with massive matrix elements (ISUB = 81, 82, 84, 85). \iteme{= 6 :} top ($\t\overline{\mathrm{t}}$) production with massive matrix elements (ISUB = 81, 82, 84, 85). \iteme{= 7 :} low ($\mathrm{l}\br{\mathrm{l}}$) production with massive matrix elements (ISUB = 81, 82, 84, 85). \iteme{= 8 :} high ($\mathrm{h}\br{\mathrm{h}}$) production with massive matrix elements (ISUB = 81, 82, 84, 85). \iteme{= 10 :} prompt photons (ISUB = 14, 18, 29). \iteme{= 11 :} $\mathrm{Z}^0$ production (ISUB = 1). \iteme{= 12 :} $\mathrm{W}^{\pm}$ production (ISUB = 2). \iteme{= 13 :} $\mathrm{Z}^0$ + jet production (ISUB = 15, 30). \iteme{= 14 :} $\mathrm{W}^{\pm}$ + jet production (ISUB = 16, 31). \iteme{= 15 :} pair production of different combinations of $\gamma$, $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$ (except $\gamma\gamma$; see \ttt{MSEL} = 10) (ISUB = 19, 20, 22, 23, 25). \iteme{= 16 :} $\H^0$ production (ISUB = 3, 102, 103, 123, 124). \iteme{= 17 :} $\H^0 \mathrm{Z}^0$ or $\H^0 \mathrm{W}^{\pm}$ (ISUB = 24, 26). \iteme{= 18 :} $\H^0$ production, combination relevant for $\e^+\e^-$ annihilation (ISUB = 24, 103, 123, 124). \iteme{= 19 :} $\H^0$, $\H'^0$ and $\mathrm{A}^0$ production, excepting pair production (ISUB = 24, 103, 123, 124, 153, 158, 171, 173, 174, 176, 178, 179). \iteme{= 21 :} $\mathrm{Z}'^0$ production (ISUB = 141). \iteme{= 22 :} $\mathrm{W}'^{\pm}$ production (ISUB = 142). \iteme{= 23 :} $\H^{\pm}$ production (ISUB = 143). \iteme{= 24 :} $\mathrm{R}^0$ production (ISUB = 144). \iteme{= 25 :} $\L_{\mathrm{Q}}$ (leptoquark) production (ISUB = 145, 162, 163, 164). \iteme{= 35:} single bottom production by $\mathrm{W}$ exchange (ISUB = 83). \iteme{= 36:} single top production by $\mathrm{W}$ exchange (ISUB = 83). \iteme{= 37:} single low production by $\mathrm{W}$ exchange (ISUB = 83). \iteme{= 38:} single high production by $\mathrm{W}$ exchange (ISUB = 83). \end{subentry} \boxsep \iteme{MSUB :}\label{p:MSUB} (D=200*0) array to be set when \ttt{MSEL=0} (for \ttt{MSEL}$\geq 1$ relevant entries are set in \ttt{PYINIT}) to choose which subset of subprocesses to include in the generation. The ordering follows the ISUB code given in section \ref{ss:ISUBcode} (with comments as given there). \begin{subentry} \iteme{MSUB(ISUB) = 0 :} the subprocess is excluded. \iteme{MSUB(ISUB) = 1 :} the subprocess is included. \itemc{Note:} when \ttt{MSEL=0}, the \ttt{MSUB} values set by the user are never changed by {\tsc{Pythia}}. If you want to combine several different `subruns', each with its own \ttt{PYINIT} call, into one single run, it is up to you to remember not only to switch on the new processes before each new \ttt{PYINIT} call, but also to switch off the old ones that are no longer desired. \end{subentry} \boxsep \iteme{KFIN(I,J) :}\label{p:KFIN} provides an option to selectively switch on and off contributions to the cross sections from the different incoming partons/particles at the hard interaction. In combination with the {\tsc{Jetset}} resonance decay switches, this also allows you to set restrictions on flavours appearing in the final state. \begin{subentry} \iteme{I :} is 1 for beam side of event and 2 for target side. \iteme{J :} enumerates flavours according to the KF code; see section \ref{ss:codes}. \iteme{KFIN(I,J) = 0 :} the parton/particle is forbidden. \iteme{KFIN(I,J) = 1 :} the parton/particle is allowed. \itemc{Note:} By default, the following are switched on: $\d$, $\u$, $\mathrm{s}$, $\c$, $\b$, $\mathrm{e}^-$, $\nu_{\mathrm{e}}$, $\mu^-$, $\nu_{\mu}$, $\tau^-$, $\nu_{\tau}$, $\mathrm{g}$, $\gamma$, $\mathrm{Z}^0$, $\mathrm{W}^+$ and their antiparticles. In particular, top is off, and has to be switched on explicitly if needed. \end{subentry} \boxsep \iteme{CKIN :}\label{p:CKIN} kinematics cuts that can be set by you before the \ttt{PYINIT} call, and that affect the region of phase space within which events are generated. Some cuts are `hardwired' while most are `softwired'. The hardwired ones are directly related to the kinematical variables used in the event selection procedure, and therefore have negligible effects on program efficiency. The most important of these are \ttt{CKIN(1) - CKIN(8)}, \ttt{CKIN(27) - CKIN(28)}, and \ttt{CKIN(31) - CKIN(32)}. The softwired ones are most of the remaining ones, that cannot be fully taken into account in the kinematical variable selection, so that generation in constrained regions of phase space may be slow. In extreme cases the phase space may be so small that the maximization procedure fails to find any allowed points at all (although some small region might still exist somewhere), and therefore switches off some subprocesses, or aborts altogether. \iteme{CKIN(1), CKIN(2) :} (D=2.,-1. GeV) range of allowed $\hat{m} = \sqrt{\hat{s}}$ values. If \ttt{CKIN(2)}$ < 0.$, the upper limit is inactive. \iteme{CKIN(3), CKIN(4) :} (D=0.,-1. GeV) range of allowed $\hat{p}_{\perp}$ values for hard $2 \to 2$ processes, with transverse momentum $\hat{p}_{\perp}$ defined in the rest frame of the hard interaction. If \ttt{CKIN(4)}$ < 0.$, the upper limit is inactive. For processes that are singular in the limit $\hat{p}_{\perp} \to 0$ (see \ttt{CKIN(6)}), \ttt{CKIN(5)} provides an additional constraint. The \ttt{CKIN(3)} and \ttt{CKIN(4)} limits can also be used in $2 \to 1 \to 2$ processes. Here, however, the product masses are not known and hence are assumed to be vanishing in the event selection. The actual $p_{\perp}$ range for massive products is thus shifted downwards with respect to the nominal one. \iteme{CKIN(5) :} (D=1. GeV) lower cut-off on $\hat{p}_{\perp}$ values, in addition to the \ttt{CKIN(3)} cut above, for processes that are singular in the limit $\hat{p}_{\perp} \to 0$ (see \ttt{CKIN(6)}). \iteme{CKIN(6) :} (D=1. GeV) hard $2 \to 2$ processes, which do not proceed only via an intermediate resonance (i.e. are $2 \to 1 \to 2$ processes), are classified as singular in the limit $\hat{p}_{\perp} \to 0$ if either or both of the two final-state products has a mass $m < $\ttt{CKIN(6)}. \iteme{CKIN(7), CKIN(8) :} (D=-10.,10.) range of allowed scattering subsystem rapidities $y = y^*$ in the c.m. frame of the event, where $y = (1/2) \ln(x_1/x_2)$. (Following the notation of this section, the variable should be given as $y^*$, but because of its frequent use, it was called $y$ in section \ref{ss:kinemtwo}.) \iteme{CKIN(9), CKIN(10) :} (D=-10.,10.) range of allowed (true) rapidities for the product with largest rapidity in a $2 \to 2$ or a $2 \to 1 \to 2$ process, defined in the c.m. frame of the event, i.e. $\max(y^*_3, y^*_4)$. Note that rapidities are counted with sign, i.e. if $y^*_3 = 1$ and $y^*_4 = -2$ then $\max(y^*_3, y^*_4) = 1$. \iteme{CKIN(11), CKIN(12) :} (D=-10.,10.) range of allowed (true) rapidities for the product with smallest rapidity in a $2 \to 2$ or a $2 \to 1 \to 2$ process, defined in the c.m. frame of the event, i.e. $\min(y^*_3, y^*_4)$. Consistency thus requires \ttt{CKIN(11)}$\leq$\ttt{CKIN(9)} and \ttt{CKIN(12)}$\leq$\ttt{CKIN(10)}. \iteme{CKIN(13), CKIN(14) :} (D=-10.,10.) range of allowed pseudorapidities for the product with largest pseudorapidity in a $2 \to 2$ or a $2 \to 1 \to 2$ process, defined in the c.m. frame of the event, i.e. $\max(\eta^*_3, \eta^*_4)$. Note that pseudorapidities are counted with sign, i.e. if $\eta^*_3 = 1$ and $\eta^*_4 = -2$ then $\max(\eta^*_3, \eta^*_4) = 1$. \iteme{CKIN(15), CKIN(16) :} (D=-10.,10.) range of allowed pseudorapidities for the product with smallest pseudorapidity in a $2 \to 2$ or a $2 \to 1 \to 2$ process, defined in the c.m. frame of the event, i.e. $\min(\eta^*_3, \eta^*_4)$. Consistency thus requires \ttt{CKIN(15)}$\leq$\ttt{CKIN(13)} and \ttt{CKIN(16)}$\leq$\ttt{CKIN(14)}. \iteme{CKIN(17), CKIN(18) :} (D=-1.,1.) range of allowed $\cos\theta^*$ values for the product with largest $\cos\theta^*$ value in a $2 \to 2$ or a $2 \to 1 \to 2$ process, defined in the c.m. frame of the event, i.e. $\max(\cos\theta^*_3,\cos\theta^*_4)$. \iteme{CKIN(19), CKIN(20) :} (D=-1.,1.) range of allowed $\cos\theta^*$ values for the product with smallest $\cos\theta^*$ value in a $2 \to 2$ or a $2 \to 1 \to 2$ process, defined in the c.m. frame of the event, i.e. $\min(\cos\theta^*_3,\cos\theta^*_4)$. Consistency thus requires \ttt{CKIN(19)}$\leq$\ttt{CKIN(17)} and \ttt{CKIN(20)}$\leq$\ttt{CKIN(18)}. \iteme{CKIN(21), CKIN(22) :} (D=0.,1.) range of allowed $x_1$ values for the parton on side 1 that enters the hard interaction. \iteme{CKIN(23), CKIN(24) :} (D=0.,1.) range of allowed $x_2$ values for the parton on side 2 that enters the hard interaction. \iteme{CKIN(25), CKIN(26) :} (D=-1.,1.) range of allowed Feynman-$x$ values, where $x_{\mrm{F}} = x_1 - x_2$. \iteme{CKIN(27), CKIN(28) :} (D=-1.,1.) range of allowed $\cos\hat{\theta}$ values in a hard $2 \to 2$ scattering, where $\hat{\theta}$ is the scattering angle in the rest frame of the hard interaction. \iteme{CKIN(31), CKIN(32) :} (D=2.,-1. GeV) range of allowed $\hat{m}' = \sqrt{\hat{s}'}$ values, where $\hat{m}'$ is the mass of the complete three- or four-body final state in $2 \to 3$ or $2 \to 4$ processes (while $\hat{m}$, constrained in \ttt{CKIN(1)} and \ttt{CKIN(2)}, here corresponds to the one- or two-body central system). If \ttt{CKIN(32)}$ < 0.$, the upper limit is inactive. \iteme{CKIN(35), CKIN(36) :} (D=0.,-1. GeV$^2$) range of allowed $|\hat{t}| = - \hat{t}$ values in $2 \to 2$ processes. Note that for deep inelastic scattering this is nothing but the $Q^2$ scale, in the limit that initial- and final-state radiation is neglected. If \ttt{CKIN(36)}$ < 0.$, the upper limit is inactive. \iteme{CKIN(37), CKIN(38) :} (D=0.,-1. GeV$^2$) range of allowed $|\hat{u}| = - \hat{u}$ values in $2 \to 2$ processes. If \ttt{CKIN(38)}$ < 0.$, the upper limit is inactive. \iteme{CKIN(39), CKIN(40) :} (D=4., -1. GeV$^2$) the $W^2$ range allowed in DIS processes, i.e. subprocess number 10. If \ttt{CKIN(40)}$ < 0.$, the upper limit is inactive. Here $W^2$ is defined in terms of $W^2 = Q^2 (1-x)/x$. This formula is not quite correct, in that \textit{(i)} it neglects the target mass (for a proton), and \textit{(ii)} it neglects initial-state photon radiation off the incoming electron. It should be good enough for loose cuts, however. \iteme{CKIN(41) - CKIN(44) :} (D=12.,-1.,12.,-1. GeV) range of allowed mass values of the two (or one) resonances produced in a `true' $2 \to 2$ process, i.e. one not (only) proceeding through a single $s$-channel resonance ($2 \to 1 \to 2$). (These are the ones listed as $2 \to 2$ in the tables in section \ref{ss:ISUBcode}.) Only particles with a width above \ttt{PARP(41)} are considered as bona fide resonances and tested against the \ttt{CKIN} limits; particles with a smaller width are put on the mass shell without applying any cuts. The exact interpretation of the \ttt{CKIN} variables depends on the flavours of the two produced resonances. \\ For two resonances like $\mathrm{Z}^0 \mathrm{W}^+$ (produced from $\mathrm{f} \mathrm{f}' \to \mathrm{Z}^0\mathrm{W}^+$), which are not identical and which are not each other's antiparticles, one has \\ \ttt{CKIN(41)}$ < m_1 < $\ttt{CKIN(42)}, and \\ \ttt{CKIN(43)}$ < m_2 < $\ttt{CKIN(44)}, \\ where $m_1$ and $m_2$ are the actually generated masses of the two resonances, and 1 and 2 are defined by the order in which they are given in the production process specification. \\ For two resonances like $\mathrm{Z}^0 \mathrm{Z}^0$, which are identical, or $\mathrm{W}^+ \mathrm{W}^-$, which are each other's antiparticles, one instead has \\ \ttt{CKIN(41)}$ < \min(m_1,m_2) < $\ttt{CKIN(42)}, and \\ \ttt{CKIN(43)}$ < \max(m_1,m_2) < $\ttt{CKIN(44)}. \\ In addition, whatever limits are set on \ttt{CKIN(1)} and, in particular, on \ttt{CKIN(2)} obviously affect the masses actually selected. \begin{subentry} \itemc{Note 1:} If \ttt{MSTP(42)=0}, so that no mass smearing is allowed, the \ttt{CKIN} values have no effect (the same as for particles with too narrow a width). \itemc{Note 2:} If \ttt{CKIN(42)}$ < $\ttt{CKIN(41)} it means that the \ttt{CKIN(42)} limit is inactive; correspondingly, if \ttt{CKIN(44)}$ < $\ttt{CKIN(43)} then \ttt{CKIN(44)} is inactive. \itemc{Note 3:} If limits are active and the resonances are identical, it is up to you to ensure that \ttt{CKIN(41)}$\leq$\ttt{CKIN(43)} and \ttt{CKIN(42)}$\leq$\ttt{CKIN(44)}. \itemc{Note 4:} For identical resonances, it is not possible to preselect which of the resonances is the lighter one; if, for instance, one $\mathrm{Z}^0$ is to decay to leptons and the other to quarks, there is no mechanism to guarantee that the lepton pair has a mass smaller than the quark one. \itemc{Note 5:} The \ttt{CKIN} values are applied to all relevant $2 \to 2$ processes equally, which may not be what one desires if several processes are generated simultaneously. Some caution is therefore urged in the use of the \ttt{CKIN(41) - CKIN(44)} values. Also in other respects, users are recommended to take proper care: if a $\mathrm{Z}^0$ is only allowed to decay into $\b\overline{\mathrm{b}}$, for example, setting its mass range to be 2--8 GeV is obviously not a good idea. \itemc{Note 6:} In principle, the machinery should work for any $2 \to 2$ process with resonances in the final state, but so far it has only been checked for processes 22--26, so also from this point some caution is urged. \end{subentry} \iteme{CKIN(45) - CKIN(48) :} (D=12.,-1.,12.,-1. GeV) range of allowed mass values of the two (or one) secondary resonances produced in a $2 \to 1 \to 2$ process (like $\mathrm{g} \mathrm{g} \to \H^0 \to \mathrm{Z}^0 \mathrm{Z}^0$) or even a $2 \to 2 \to 4$ (or 3) process (like $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{Z}^0 \H^0 \to \mathrm{Z}^0 \mathrm{W}^+ \mathrm{W}^-$). Note that these \ttt{CKIN} values only affect the secondary resonances; the primary ones are constrained by \ttt{CKIN(1)}, \ttt{CKIN(2)} and \ttt{CKIN(41) - CKIN(44)} (indirectly, of course, the choice of primary resonance masses affects the allowed mass range for the secondary ones). What is considered to be a resonance is defined by \ttt{PARP(41)}; particles with a width smaller than this are automatically put on the mass shell. The description closely parallels the one given for \ttt{CKIN(41) - CKIN(44)}. Thus, for two resonances that are not identical or each other's antiparticles, one has \\ \ttt{CKIN(45)}$ < m_1 < $\ttt{CKIN(46)}, and \\ \ttt{CKIN(47)}$ < m_2 < $\ttt{CKIN(48)}, \\ where $m_1$ and $m_2$ are the actually generated masses of the two resonances, and 1 and 2 are defined by the order in which they given in the decay channel specification in the program (see e.g. output from \ttt{PYSTAT(2)} or \ttt{LULIST(12)}). For two resonances that are identical or each other's antiparticles, one instead has \\ \ttt{CKIN(45)}$ < \min(m_1,m_2) < $\ttt{CKIN(46)}, and \\ \ttt{CKIN(47)}$ < \max(m_1,m_2) < $\ttt{CKIN(48)}. \begin{subentry} \itemc{Notes 1 - 5:} as for \ttt{CKIN(41) - CKIN(44)}, with trivial modifications. \itemc{Note 6:} Setting limits on secondary resonance masses is possible in any of the channels of the allowed types (see above). However, so far only $\H^0 \to \mathrm{Z}^0 \mathrm{Z}^0$ and $\H^0 \to \mathrm{W}^+ \mathrm{W}^-$ have been fully implemented, such that an arbitrary mass range below the na\"{\i}ve mass threshold may be picked. For other possible resonances, any restrictions made on the allowed mass range are not reflected in the cross section; and further it is not recommendable to pick mass windows that make a decay on the mass shell impossible. These limitations will be relaxed in future versions. \end{subentry} \iteme{CKIN(51) - CKIN(56) :} (D=0.,-1.,0.,-1.,0.,-1. GeV) range of allowed transverse momenta in a true $2 \to 3$ process. Currently two different alternatives are around. For subprocess 131, the $p_{\perp}$ of the first product (the $\mathrm{Z}^0$) is set by \ttt{CKIN(3)} and \ttt{CKIN(4)}, while for the quark and antiquark $p_{\perp}$'s one has \\ \ttt{CKIN(51)}$<\min(p_{\perp \mathrm{q}},p_{\perp \overline{\mathrm{q}}})<$\ttt{CKIN(52)}, and \\ \ttt{CKIN(53)}$<\max(p_{\perp \mathrm{q}},p_{\perp \overline{\mathrm{q}}})<$\ttt{CKIN(54)}. \\ Negative \ttt{CKIN(52)} and \ttt{CKIN(54)} values means that the corresponding limits are inactive. For subprocesses 121--124, and their $\H'^0$ and $\mathrm{A}^0$ equivalents (173, 174, 178, 179, 181, 182, 186, 187), \ttt{CKIN(51) - CKIN(54)} again corresponds to $p_{\perp}$ ranges for scattered partons, but in order of appearance, i.e. \ttt{CKIN(51) - CKIN(52)} for the parton scattered off the beam and \ttt{CKIN(53) - CKIN(54)} for the one scattered off the target. \ttt{CKIN(55)} and \ttt{CKIN(56)} here sets $p_{\perp}$ limits for the third product, the $\H^0$, i.e. the \ttt{CKIN(3)} and \ttt{CKIN(4)} values have no effect for this process. Since the $p_{\perp}$ of the Higgs is not one of the primary variables selected, any constraints here may mean reduced Monte Carlo efficiency, while for these processes \ttt{CKIN(51) - CKIN(54)} are `hardwired' and therefore do not cost anything. \end{entry} \subsection{The General Switches and Parameters} \label{ss:PYswitchpar} The \ttt{PYPARS} common block contains the status code and parameters that regulate the performance of the program. All of them are provided with sensible default values, so that a novice user can neglect them, and only gradually explore the full range of possibilities. Some of the switches and parameters in \ttt{PYPARS} will be described later, in the shower and beam remnants sections. \drawbox{COMMON/PYPARS/MSTP(200),PARP(200),MSTI(200),PARI(200)}% \label{p:PYPARS1} \begin{entry} \itemc{Purpose:} to give access to status code and parameters that regulate the performance of the program. If the default values, denoted below by (D=\ldots), are not satisfactory, they must in general be changed before the \ttt{PYINIT} call. Exceptions, i.e. variables that can be changed for each new event, are denoted by (C). \iteme{MSTP(1) :}\label{p:MSTP} (D=3) maximum number of generations. Automatically set $\leq 4$. \iteme{MSTP(2) :} (D=1) calculation of $\alpha_{\mathrm{s}}$ at hard interaction, in the routine \ttt{ULALPS}. \begin{subentry} \iteme{= 0 :} $\alpha_{\mathrm{s}}$ is fixed at value \ttt{PARU(111)}. \iteme{= 1 :} first-order running $\alpha_{\mathrm{s}}$. \iteme{= 2 :} second-order running $\alpha_{\mathrm{s}}$. \end{subentry} \iteme{MSTP(3) :} (D=2) selection of $\Lambda$ value in $\alpha_{\mathrm{s}}$ for \ttt{MSTP(2)}$\geq 1$. \begin{subentry} \iteme{= 1 :} $\Lambda$ is given by \ttt{PARP(1)} for hard interactions, by \ttt{PARP(61)} for space-like showers, by \ttt{PARP(72)} for time-like showers not from a resonance decay, and by \ttt{PARJ(81)} for time-like ones from a resonance decay (including e.g. $\gamma/\mathrm{Z}^0 \to \mathrm{q}\overline{\mathrm{q}}$ decays, i.e. conventional $\e^+\e^-$ physics). This $\Lambda$ is assumed to be valid for 5 flavours; for the hard interaction the number of flavours assumed can be changed by \ttt{MSTU(112)}. \iteme{= 2 :} $\Lambda$ value is chosen according to the parton-distribution-function para\-metri\-zations, i.e. $\Lambda$ = 0.20 GeV for EHLQ1, = 0.29 GeV for EHLQ2, = 0.20 GeV for DO1, = 0.40 GeV for DO2, = 0.213 GeV for CTEQ2M, = 0.208 GeV for CTEQ2MS, = 0.208 GeV for CTEQ2MF, = 0.322 GeV for CTEQ2ML, = 0.190 GeV for CTEQ2L, = 0.235 GeV for CTEQ2D, = 0.25 GeV for GRV LO, and similarly for parton-distribution functions in the \tsc{Pdflib} library (cf. (\ttt{MSTP(51)}, \ttt{MSTP(52)}). The choice is always based on the proton parton-distribution set selected, i.e. is unaffected by pion and photon parton-distribution selection. All the $\Lambda$ values above are assumed to refer to 4 flavours, and \ttt{MSTU(112)} is set accordingly. This $\Lambda$ value is used both for the hard scattering and the initial- and final-state radiation. The ambiguity in the choice of the $Q^2$ argument still remains (see \ttt{MSTP(32)}, \ttt{MSTP(64)} and \ttt{MSTJ(44)}). This $\Lambda$ value is used also for \ttt{MSTP(57)=0}, but the sensible choice here would be to use \ttt{MSTP(2)=0} and have no initial- or final-state radiation. This option does \textit{not} change the \ttt{PARJ(81)} value of timelike parton showers in resonance decays, so that LEP experience on this specific parameter is not overwritten unwittingly. Therefore \ttt{PARJ(81)} can be updated completely independently. \iteme{= 3 :} as \ttt{=2}, except that here also \ttt{PARJ(81)} is overwritten in accordance with the $\Lambda$ value of the proton parton-distribution-function set. \end{subentry} \iteme{MSTP(4) :} (D=0) treatment of the Higgs sector, predominantly the neutral one. \begin{subentry} \iteme{= 0 :} the $\H^0$ is given the Standard Model Higgs couplings, while $\H'^0$ and $\mathrm{A}^0$ couplings should be set by the user in \ttt{PARU(171) - PARU(175)} and \ttt{PARU(181) - PARU(185)}, respectively. \iteme{= 1 :} the user should set couplings for all three Higgses, for the $\H^0$ in \ttt{PARU(161) - PARU(165)}, and for the $\H'^0$ and $\mathrm{A}^0$ as above. \iteme{= 2 :} the mass of $\H^0$ in \ttt{PMAS(25,1)} and the $\tan\beta$ value in \ttt{PARU(141)} are used to derive $\H'^0$, $\mathrm{A}^0$ and $\H^{\pm}$ masses, and $\H^0$, $\H'^0$, $\mathrm{A}^0$ and $\H^{\pm}$ couplings, using the relations of the Minimal Supersymmetric extension of the Standard Model at Born level \cite{Gun90}. Existing masses and couplings are overwritten by the derived values. See section \ref{sss:extneutHclass} for discussion on parameter constraints. \iteme{= 3:} as \ttt{=2}, but using relations at the one-loop level. This option is not yet implemented. \end{subentry} \iteme{MSTP(5) :} (D=0) presence of anomalous couplings in processes. \begin{subentry} \iteme{= 0 :} absent. \iteme{$\geq$1 :} present, wherever implemented. See section \ref{sss:ancoupclass} for further details. \end{subentry} \iteme{MSTP(6) :} (D=0) usage of the fourth-generation fermions to simulate other fermion kinds. \begin{subentry} \iteme{= 0 :} none, i.e. can be used as a standard fourth generation. \iteme{= 1 :} excited fermions, as present in compositeness scenarios; see section \ref{sss:qlstarclass}. \end{subentry} \iteme{MSTP(7) :} (D=0) choice of heavy flavour in subprocesses 81--85. Does not apply for \ttt{MSEL=4-8}, where the MSEL value always takes precedence. \begin{subentry} \iteme{= 0 :} for processes 81--84 (85) the `heaviest' flavour allowed for gluon (photon) splitting into a quark--antiquark (fermion--antifermion) pair, as set in the \ttt{MDME} array. Note that `heavy' is defined as the one with largest KF code, so that leptons take precedence if they are allowed. \iteme{= 1 - 8 :} pick this particular quark flavour; e.g., \ttt{MSTP(7)=6} means that top will be produced. \iteme{= 11 - 18 :} pick this particular lepton flavour. Note that neutrinos are not possible, i.e. only 11, 13, 15 and 17 are meaningful alternatives. Lepton pair production can only occur in process 85, so if any of the other processes have been switched on they are generated with the same flavour as would be obtained in the option \ttt{MSTP(7)=0}. \end{subentry} \iteme{MSTP(8) :} (D=0) choice of electroweak parameters to use in the decay widths of resonances ($\mathrm{W}$, $\mathrm{Z}$, $\H$, \ldots) and cross sections (production of $\mathrm{W}$'s, $\mathrm{Z}$'s, $\H$'s, \ldots). \begin{subentry} \iteme{= 0 :} everything is expressed in terms of a running $\alpha_{\mathrm{em}}(Q^2)$ and a fixed $\sin^2 \! \theta_W$, i.e. $G_{\mathrm{F}}$ is nowhere used. \iteme{= 1 :} a replacement is made according to $\alpha_{\mathrm{em}}(Q^2) \to \sqrt{2} G_{\mathrm{F}} m_{\mathrm{W}}^2 \sin^2 \! \theta_W / \pi$ in all widths and cross sections. If $G_{\mathrm{F}}$ and $m_{\mathrm{Z}}$ are considered as given, this means that $\sin^2 \! \theta_W$ and $m_{\mathrm{W}}$ are the only free electroweak parameter. \iteme{= 2 :} a replacement is made as for \ttt{=1}, but additionally $\sin^2 \! \theta_W$ is constrained by the relation $\sin^2 \! \theta_W = 1 - m_{\mathrm{W}}^2/m_{\mathrm{Z}}^2$. This means that $m_{\mathrm{W}}$ remains as a free parameter, but that the $\sin^2 \! \theta_W$ value in \ttt{PARU(102)} is never used, \textit{except} in the vector couplings in the combination $v = a - 4 \sin^2 \! \theta_W e$. This latter degree of freedom enters e.g. for forward-backward asymmetries in $\mathrm{Z}^0$ decays. \itemc{Note:} This option does not affect the emission of real photons in the initial and final state, where $\alpha_{\mathrm{em}}$ is always used. However, it does affect also purely electromagnetic hard processes, such as $\mathrm{q} \overline{\mathrm{q}} \to \gamma \gamma$. \end{subentry} \iteme{MSTP(11) :} (D=1) use of electron parton distribution in $\e^+\e^-$ and $\e\p$ interactions. \begin{subentry} \iteme{= 0 :} no, i.e. electron carries the whole beam energy. \iteme{= 1 :} yes, i.e. electron carries only a fraction of beam energy in agreement with next-to-leading electron parton-distribution function, thereby including the effects of initial-state bremsstrahlung. \end{subentry} \iteme{MSTP(12) :} (D=0) use of $\mathrm{e}^-$ (`sea', i.e. from $\mathrm{e} \to \gamma \to \mathrm{e}$), $\mathrm{e}^+$, quark and gluon distribution functions inside an electron. \begin{subentry} \iteme{= 0 :} off. \iteme{= 1 :} on, provided that \ttt{MSTP(11)}$\geq 1$. Quark and gluons distributions are obtained by numerical convolution of the photon content inside an electron (as given by the bremsstrahlung spectrum of \ttt{MSTP(11)=1}) with the quark and gluon content inside a photon. The required numerical precision is set by \ttt{PARP(14)}. Since the need for numerical integration makes this option somewhat more time-consuming than ordinary parton-distribution evaluation, one should only use it when studying processes where it is needed. \end{subentry} \iteme{MSTP(13) :} (D=1) choice of $Q^2$ range over which electrons are assumed to radiate photons; affects normalization of $\mathrm{e}^-$ (sea), $\mathrm{e}^+$, $\gamma$, quark and gluon distributions inside an electron. \begin{subentry} \iteme{= 1 :} range set by $Q^2$ argument of parton-distribution-function call, i.e. by $Q^2$ scale of the hard interaction. Therefore parton distributions are proportional to $\ln(Q^2/m_e^2)$. This is normally most appropriate for $\e^+\e^-$ annihilation. \iteme{= 2 :} range set by the user-determined $Q_{\mathrm{max}}^2$, given in \ttt{PARP(13)}. Parton distributions are assumed to be proportional to $\ln((Q_{\mathrm{max}}^2/m_e^2)(1-x)/x^2)$. This is normally most appropriate for photoproduction, where the electron is supposed to go undetected, i.e. scatter less than $Q_{\mathrm{max}}^2$. \itemc{Note:} the choice of effective range is especially touchy for the quark and gluon distributions. An (almost) on-the-mass-shell photon has a VMD piece that dies away for a virtual photon. A simple folding of distribution functions does not take this into account properly. Therefore the contribution from $Q$ values above the $\rho$ mass should be suppressed. A choice of $Q_{\mathrm{max}} \approx 1$~GeV is then appropriate for a photoproduction limit description of physics. \end{subentry} \iteme{MSTP(14) :} (D=0) structure of incoming photon beam or target (does not affect photon inside electron, only photons appearing as argument in the \ttt{PYINIT} call). \begin{subentry} \iteme{= 0 :} a photon is assumed to be point-like (a direct photon), i.e. can only interact in processes which explicitly contain the incoming photon, such as $\mathrm{f}_i \gamma \to \mathrm{f}_i \mathrm{g}$ for $\gamma\mathrm{p}$ interactions. In $\gamma\gamma$ interactions both photons are direct, i.e the main process is $\gamma \gamma \to \mathrm{f}_i \overline{\mathrm{f}}_i$. \iteme{= 1 :} a photon is assumed to be resolved, i.e. can only interact through its constituent quarks and gluons, giving either high-$p_{\perp}$ parton--parton scatterings or low-$p_{\perp}$ events. Hard processes are calculated with the use of the full photon parton distributions. In $\gamma\gamma$ interactions both photons are resolved. \iteme{= 2 :} a photon is assumed resolved, but only the VMD piece is included in the parton distributions, which therefore mainly are scaled-down versions of the $\rho^0 / \pi^0$ ones. Both high-$p_{\perp}$ parton--parton scatterings and low-pT events are allowed. In $\gamma\gamma$ interactions both photons are VMD-like. \iteme{= 3 :} a photon is assumed resolved, but only the anomalous piece of the photon parton distributions is included. Only high-$p_{\perp}$ parton--parton scatterings are allowed. In $\gamma\gamma$ interactions both photons are anomalous. \iteme {= 4 :} in $\gamma\gamma$ interactions one photon is direct and the other resolved. A typical process is thus $\mathrm{f}_i \gamma \to \mathrm{f}_i \mathrm{g}$. Hard processes are calculated with the use of the full photon parton distributions for the resolved photon. Both possibilities of which photon is direct are included, in event topologies and in cross sections. This option cannot be used in configurations with only one incoming photon. \iteme{= 5 :} in $\gamma\gamma$ interactions one photon is direct and the other VMD-like. Both possibilities of which photon is direct are included, in event topologies and in cross sections. This option cannot be used in configurations with only one incoming photon. \iteme{= 6 :} in $\gamma\gamma$ interactions one photon is direct and the other anomalous. Both possibilities of which photon is direct are included, in event topologies and in cross sections. This option cannot be used in configurations with only one incoming photon. \iteme{= 7 :} in $\gamma\gamma$ interactions one photon is VMD-like and the other anomalous. Only high-$p_{\perp}$ parton--parton scatterings are allowed. Both possibilities of which photon is VMD-like are included, in event topologies and in cross sections. This option cannot be used in configurations with only one incoming photon. \iteme{= 10 :} the VMD, direct and anomalous components of the photon are automatically mixed. For $\gamma\mathrm{p}$ interactions, this means an automatic mixture of the three classes 0, 2 and 3 above \cite{Sch93,Sch93a}, for $\gamma\gamma$ ones a mixture of the six classes 0, 2, 3, 5, 6 and 7 above \cite{Sch94a}. Various restrictions exist for this option, as discussed in section \ref{sss:photoprodclass}. \itemc{Note:} our best understanding of how to mix event classes is provided by the option 10 above, which also can be obtained by combining three (for $\gamma\mathrm{p}$) or six (for $\gamma\gamma$) separate runs. In a simpler alternative the VMD and anomalous classes are joined into a single resolved class. Then $\gamma\mathrm{p}$ physics only requires two separate runs, with 0 and 1, and $\gamma\gamma$ physics requires three, with 0, 1 and 4. \end{subentry} \iteme{MSTP(15) :} (D=5) possibility to modify the nature of the anomalous photon component (as used with the appropriate \ttt{MSTP(14)} options), in particular with respect to the scale choices and cut-offs of hard processes. This option is mainly intended for comparative studies and should not normally be touched. \begin{subentry} \iteme{= 0 :} none, i.e. the same treatment as for the VMD component. \iteme{= 1 :} evaluate the anomalous parton distributions at a scale $Q^2/$\ttt{PARP(17)}$^2$. \iteme{= 2 :} as \ttt{=1}, but instead of \ttt{PARP(17)} use either \ttt{PARP(81)/PARP(15)} or \ttt{PARP(82)/PARP(15)}, depending on \ttt{MSTP(82)} value. \iteme{= 3 :} evaluate the anomalous parton distribution functions of the photon as $f^{\gamma,\mrm{anom}}(x, Q^2, p_0^2) - f^{\gamma,\mrm{anom}}(x, Q^2, r^2 Q^2)$ with $r = $\ttt{PARP(17)}. \iteme{= 4 :} as \ttt{=3}, but instead of \ttt{PARP(17)} use either \ttt{PARP(81)/PARP(15)} or \ttt{PARP(82)/PARP(15)}, depending on \ttt{MSTP(82)} value. \iteme{= 5 :} use larger $p_{\perp\mathrm{min}}$ for the anomalous component than for the VMD one, but otherwise no difference. \end{subentry} \iteme{MSTP(21) :} (D=1) nature of fermion--fermion scatterings simulated in process 10 by $t$-channel exchange. \begin{subentry} \iteme{= 0 :} all off. \iteme{= 1 :} full mixture of $\gamma^* / \Z^0$ neutral current and $\mathrm{W}^{\pm}$ charged current. \iteme{= 2 :} $\gamma$ neutral current only. \iteme{= 3 :} $\mathrm{Z}^0$ neutral current only. \iteme{= 4 :} $\gamma^* / \Z^0$ neutral current only. \iteme{= 5 :} $\mathrm{W}^{\pm}$ charged current only. \end{subentry} \iteme{MSTP(22) :} (D=0) special override of normal $Q^2$ definition used for maximum of parton-shower evolution, intended for deep inelastic scattering) in lepton--hadron events, see section \ref{ss:showrout}. \iteme{MSTP(23) :} (D=1) for deep inelastic scattering processes (10 and 83), this option allows the $x$ and $Q^2$ of the original hard scattering to be retained by the final electron. \begin{subentry} \iteme{= 0 :} no correction procedure, i.e. $x$ and $Q^2$ of the scattered electron differ from the originally generated $x$ and $Q^2$. \iteme{= 1 :} post facto correction, i.e. the change of electron momentum, by initial and final QCD radiation, primordial $k_{\perp}$ and beam remnant treatment, is corrected for by a shuffling of momentum between the electron and hadron side in the final state. Only process 10 is corrected, while process 83 is not. \iteme{= 2 :} as \ttt{=1}, except that both process 10 and 83 are treated. This option is dangerous, especially for top, since it may well be impossible to `correct' in process 83: the standard DIS kinematics definitions are based on the assumption of massless quarks. Therefore infinite loops are not excluded. \itemc{Note:} the correction procedure will fail for a fraction of the events, which are thus rejected (and new ones generated in their place). The correction option is not unambiguous, and should not be taken too seriously. For very small $Q^2$ values, the $x$ is not exactly preserved even after this procedure. \end{subentry} \iteme{MSTP(31) :} (D=1) parametrization of total, elastic and diffractive cross sections. \begin{subentry} \iteme{= 0 :} everything is to be set by you yourself in the \ttt{PYINT7} common block. For photoproduction, additionally you need to set \ttt{VINT(281)}. Normally you would set these values once and for all before the \ttt{PYINIT} call, but if you run with variable energies (see \ttt{MSTP(171)}) you can also set it before each new \ttt{PYEVNT} call. \iteme{= 1 :} Donnachie--Landshoff for total cross section \cite{Don92}, and Schuler--Sj\"ostrand for elastic and diffractive cross sections \cite{Sch94,Sch93a}. \end{subentry} \iteme{MSTP(32) :} (D=2) $Q^2$ definition in hard scattering for $2 \to 2$ processes. For resonance production $Q^2$ is always chosen to be $\hat{s} = m_R^2$, where $m_R$ is the mass of the resonance. For gauge boson scattering processes $VV \to VV$ the $\mathrm{W}$ or $\mathrm{Z}^0$ squared mass is used as scale in parton distributions. See \ttt{PARP(34)} for a possibility to modify the choice below by a multiplicative factor. \begin{subentry} \iteme{= 1 :} $Q^2 = 2 \hat{s} \hat{t} \hat{u} / (\hat{s}^2 + \hat{t}^2 + \hat{u}^2)$. \iteme{= 2 :} $Q^2 = (m_{\perp 1}^2 + m_{\perp 2}^2)/2$. \iteme{= 3 :} $Q^2 = \min(-\hat{t}, -\hat{u})$. \iteme{= 4 :} $Q^2 = \hat{s}$. \iteme{= 5 :} $Q^2 = -\hat{t}$. \end{subentry} \iteme{MSTP(33) :} (D=0) inclusion of $K$ factors in hard cross sections for parton--parton interactions (i.e. for incoming quarks and gluons). \begin{subentry} \iteme{= 0 :} none, i.e. $K = 1$. \iteme{= 1 :} a common $K$ factor is used, as stored in \ttt{PARP(31)}. \iteme{= 2 :} separate factors are used for ordinary (\ttt{PARP(31)}) and colour annihilation graphs (\ttt{PARP(32)}). \iteme{= 3 :} A $K$ factor is introduced by a shift in the $\alpha_{\mathrm{s}}$ $Q^2$ argument, $\alpha_{\mathrm{s}} = \alpha_{\mathrm{s}}($\ttt{PARP(33)}$Q^2)$. \end{subentry} \iteme{MSTP(34) :} (D=1) use of interference term in matrix elements for QCD processes, see section \ref{sss:QCDjetclass}. \begin{subentry} \iteme{= 0 :} excluded (i.e. string-inspired matrix elements). \iteme{= 1 :} included (i.e. conventional QCD matrix elements). \itemc{Note:} for the option \ttt{MSTP(34)=1}, i.e. interference terms included, these terms are divided between the different possible colour configurations according to the pole structure of the (string-inspired) matrix elements for the different colour configurations. \end{subentry} \iteme{MSTP(35) :} (D=0) threshold behaviour for heavy-flavour production, i.e. ISUB = 81, 82, 84, 85, and also for $\mathrm{Z}$ and $\mathrm{Z}'$ decays. The non-standard options are mainly intended for top, but can be used, with less theoretical reliability, also for charm and bottom (for $\mathrm{Z}$ and $\mathrm{Z}'$ only top and heavier flavours are affected). The threshold factors are given in eqs.~(\ref{pp:threshenh}) and (\ref{pp:threshsup}). \begin{subentry} \iteme{= 0 :} na\"{\i}ve lowest-order matrix-element behaviour. \iteme{= 1 :} enhancement or suppression close to threshold, according to the colour structure of the process. The $\alpha_{\mathrm{s}}$ value appearing in the threshold factor (which is not the same as the $\alpha_{\mathrm{s}}$ of the lowest-order $2 \to 2$ process) is taken to be fixed at the value given in \ttt{PARP(35)}. The threshold factor used in an event is stored in \ttt{PARI(81)}. \iteme{= 2 :} as \ttt{=1}, but the $\alpha_{\mathrm{s}}$ value appearing in the threshold factor is taken to be running, with argument $Q^2 = m_{\mathrm{Q}} \sqrt{ (\hat{m} - 2m_{\mathrm{Q}})^2 + \Gamma_{\mathrm{Q}}^2}$. Here $m_{\mathrm{Q}}$ is the nominal heavy-quark mass, $\Gamma_{\mathrm{Q}}$ is the width of the heavy-quark-mass distribution, and $\hat{m}$ is the invariant mass of the heavy-quark pair. The $\Gamma_{\mathrm{Q}}$ value has to be stored by the user in \ttt{PARP(35)}. The regularization of $\alpha_{\mathrm{s}}$ at low $Q^2$ is given by \ttt{MSTP(36)}. \end{subentry} \iteme{MSTP(36) :} (D=2) regularization of $\alpha_{\mathrm{s}}$ in the limit $Q^2 \to 0$ for the threshold factor obtainable in the \ttt{MSTP(35)=2} option; see \ttt{MSTU(115)} for a list of the possibilities. \iteme{MSTP(37) :} (D=1) inclusion of running quark masses in Higgs production ($\mathrm{q} \overline{\mathrm{q}} \to \H^0$) and decay ($\H^0 \to \mathrm{q} \overline{\mathrm{q}}$) couplings. Also included for charged Higgs production and decay, but there only for the down-type quark, since the up-type one normally is a top quark, with $m_{\t} \approx m_{\H}$. \begin{subentry} \iteme{= 0 :} not included, i.e. fixed quark masses are used according to the values in the \ttt{PMAS} array. \iteme{= 1 :} included, with running starting from the value given in the \ttt{PMAS} array, at a $Q_0$ scale of \ttt{PARP(37)} times the quark mass itself, up to a $Q$ scale given by the Higgs mass. This option only works when $\alpha_{\mathrm{s}}$ is allowed to run (so one can define a $\Lambda$ value). Therefore it is only applied if additionally \ttt{MSTP(2)}$\geq 1$. \end{subentry} \iteme{MSTP(38) :} (D=5) handling of quark loop masses in the box graphs $\mathrm{g} \mathrm{g} \to \gamma \gamma$ and $\mathrm{g} \mathrm{g} \to \mathrm{g} \gamma$. \begin{subentry} \iteme{= 0 :} the program will for each flavour automatically choose the massless approximation for light quarks and the full massive formulae for heavy quarks, with a dividing line between light and heavy quarks that depends on the actual $\hat{s}$ scale. \iteme{$\geq$1 :} the program will use the massless approximation throughout, assuming the presence of \ttt{MSTP(38)} effectively massless quark species (however, at most 8). Normally one would use \ttt{=5} for the inclusion of all quarks up to bottom, and \ttt{=6} to include top as well. \itemc{Warning:} for \ttt{=0}, numerical instabilities may arise for scattering at small angles. Users are therefore recommended in this case to set \ttt{CKIN(27)} and \ttt{CKIN(28)} so as to exclude the range of scattering angles that are not of interest anyway. \end{subentry} \iteme{MSTP(39) :} (D=2) choice of $Q^2$ scale for parton distributions and initial state parton showers in processes $\mathrm{g}\g$ or $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{Q} \overline{\mathrm{Q}} \H$. \begin{subentry} \iteme{= 1 :} $m_{\mathrm{Q}}^2$. \iteme{= 2 :} $\max(m_{\perp\mathrm{Q}}^2,m_{\perp\overline{\mathrm{Q}}}^2 ) = m_{\mathrm{Q}}^2 + \max(p_{\perp\mathrm{Q}}^2 , p_{\perp\overline{\mathrm{Q}}}^2)$. \iteme{= 3 :} $m_{\H}^2$. \iteme{= 4 :} $\hat{s} = (p_{\H} + p_{\mathrm{Q}} + p_{\overline{\mathrm{Q}}})^2$. \end{subentry} \iteme{MSTP(41) :} (D=1) master switch for all resonance decays ($\mathrm{Z}^0$, $\mathrm{W}^{\pm}$, $\H^0$, $\mathrm{Z}'^0$, $\mathrm{W}'^{\pm}$, $\H'^0$, $\mathrm{A}^0$, $\H^{\pm}$, $\L_{\mathrm{Q}}$, $\mathrm{R}^0$, $\d^*$, $\u^*$, \ldots). \begin{subentry} \iteme{= 0 :} off. \iteme{= 1 :} on. \itemc{Note:} also for \ttt{MSTP(41)=1} it is possible to switch off the decays of specific resonances by using the \ttt{MDCY(KC,1)} switches in {\tsc{Jetset}}. However, since the \ttt{MDCY} values are overwritten in the \ttt{PYINIT} call, individual resonances should be switched off after the \ttt{PYINIT} call. \itemc{Warning:} leptoquark decays must not be switched off if one later on intends to let leptoquarks decay (with \ttt{LUEXEC}); see section \ref{sss:LQclass}. \end{subentry} \iteme{MSTP(42) :} (D=1) mass treatment in $2 \to 2$ processes, where the final-state resonances have finite width (see \ttt{PARP(41)}). (Does not apply for the production of a single $s$-channel resonance, where the mass appears explicitly in the cross section of the process, and thus is always selected with width.) \begin{subentry} \iteme{= 0 :} particles are put on the mass shell. \iteme{= 1 :} mass generated according to a Breit--Wigner. \end{subentry} \iteme{MSTP(43) :} (D=3) treatment of $\mathrm{Z}^0/\gamma^*$ interference in matrix elements. So far implemented in subprocesses 1, 15, 19, 22, 30 and 35; in other processes what is called a $\mathrm{Z}^0$ is really a $\mathrm{Z}^0$ only, without the $\gamma^*$ piece. \begin{subentry} \iteme{= 1 :} only $\gamma^*$ included. \iteme{= 2 :} only $\mathrm{Z}^0$ included. \iteme{= 3 :} complete $\mathrm{Z}^0/\gamma^*$ structure (with interference) included. \end{subentry} \iteme{MSTP(44) :} (D=7) treatment of $\mathrm{Z}'^0/\mathrm{Z}^0/\gamma^*$ interference in matrix elements. \begin{subentry} \iteme{= 1 :} only $\gamma^*$ included. \iteme{= 2 :} only $\mathrm{Z}^0$ included. \iteme{= 3 :} only $\mathrm{Z}'^0$ included. \iteme{= 4 :} only $\mathrm{Z}^0/\gamma^*$ (with interference) included. \iteme{= 5 :} only $\mathrm{Z}'^0/\gamma^*$ (with interference) included. \iteme{= 6 :} only $\mathrm{Z}'^0/\mathrm{Z}^0$ (with interference) included. \iteme{= 7 :} complete $\mathrm{Z}'^0/\mathrm{Z}^0/\gamma^*$ structure (with interference) included. \end{subentry} \iteme{MSTP(45) :} (D=3) treatment of $\mathrm{W}\W \to \mathrm{W}\W$ structure (ISUB = 77). \begin{subentry} \iteme{= 1 :} only $\mathrm{W}^+\mathrm{W}^+ \to \mathrm{W}^+\mathrm{W}^+$ and $\mathrm{W}^-\mathrm{W}^- \to \mathrm{W}^-\mathrm{W}^-$ included. \iteme{= 2 :} only $\mathrm{W}^+\mathrm{W}^- \to \mathrm{W}^+\mathrm{W}^-$ included. \iteme{= 3 :} all charge combinations $\mathrm{W}\W \to \mathrm{W}\W$ included. \end{subentry} \iteme{MSTP(46) :} (D=1) treatment of $VV \to V'V'$ structures (ISUB = 71--77), where $V$ represents a longitudinal gauge boson. \begin{subentry} \iteme{= 0 :} only $s$-channel Higgs exchange included (where existing). With this option, subprocesses 71--72 and 76--77 will essentially be equivalent to subprocesses 5 and 8, respectively, with the proper decay channels (i.e. only $\mathrm{Z}^0\mathrm{Z}^0$ or $\mathrm{W}^+\mathrm{W}^-$) set via \ttt{MDME}. The description obtained for subprocesses 5 and 8 in this case is more sophisticated, however; see section \ref{sss:heavySMHclass}. \iteme{= 1 :} all graphs contributing to $VV \to V'V'$ processes are included. \iteme{= 2 :} only graphs not involving Higgs exchange (either in $s$, $t$ or $u$ channel) are included; this option then gives the na\"{\i}ve behaviour one would expect if no Higgs exists, including unphysical unitarity violations at high energies. \iteme{= 3 :} the strongly interacting Higgs-like model of Dobado, Herrero and Terron \cite{Dob91} with Pad\'e unitarization. Note that to use this option it is necessary to set the Higgs mass to a large number like 20 TeV (i.e. \ttt{PMAS(25,1)=20000}). The parameter $\nu$ is stored in \ttt{PARP(44)}, but should not have to be changed. \iteme{= 4 :} as \ttt{=3}, but with K-matrix unitarization. \iteme{= 5 :} the strongly interacting QCD-like model of Dobado, Herrero and Terron \cite{Dob91} with Pad\'e unitarization. The parameter $\nu$ is stored in \ttt{PARP(44)}, but should not have to be changed. The effective techni-$\rho$ mass in this model is stored in \ttt{PARP(45)}; by default it is 2054 GeV, which is the expected value for three technicolors, based on scaling up the ordinary $\rho$ mass appropriately. \iteme{= 6 :} as \ttt{=5}, but with K-matrix unitarization. While \ttt{PARP(45)} still is a parameter of the model, this type of unitarization does not give rise to a resonance at a mass of \ttt{PARP(45)}. \end{subentry} \iteme{MSTP(47) :} (D=1) (C) angular orientation of decay products of resonances ($\mathrm{Z}^0$, $\mathrm{W}^{\pm}$, $\H^0$, $\mathrm{Z}'^0$, $\mathrm{W}'^{\pm}$, etc.), either when produced singly or in pairs (also from an $\H^0$ decay), or in combination with a single quark, gluon or photon. \begin{subentry} \iteme{= 0 :} independent decay of each resonance, isotropic in c.m. frame of the resonance. \iteme{= 1 :} correlated decay angular distributions according to proper matrix elements, to the extent these are known. \end{subentry} \iteme{MSTP(48) :} (D=2) (C) possibility to switch between top decay before or after fragmentation. As a rule of thumb, option 0 is recommendable for top masses below 120 GeV and option 1 above that, but clearly there is a gradual transition between the two. \begin{subentry} \iteme{= 0 :} top quarks fragment to top hadrons, which subsequently decay. The $\b$ quark may be allowed to shower, see \ttt{MSTJ(27)}. The $\mathrm{W}$ produced in the decay is code 89. \iteme{= 1 :} top quarks decay, $\t \to \b \mathrm{W}^+$, and thereafter the $\b$ quark fragments. Parton showering of the $\b$ is automatically included, but can be switched off with \ttt{MSTP(71)}. The $\mathrm{W}$ has the ordinary code 24, and is allowed to decay isotropically. \iteme{= 2 :} as \ttt{=1}, except that the $\mathrm{W}$ decay is anisotropic, as expected from $\mathrm{W}$ polarization in the top decay. \itemc{Note:} in options 1 and 2 the cross section is reduced to take into account restrictions on allowed decay modes, while no such reduction occurs for option 0. See further section \ref{sss:heavflavclass}. \end{subentry} \iteme{MSTP(49) :} (D=2) (C) possibility to switch between fourth generation decay before or after fragmentation. For the quarks $\mathrm{h}$ and $\mathrm{l}$ the meaning is exactly as \ttt{MSTP(48)} is for the $\t$ quark. For the lepton $\chi$ the difference is whether decay is handled as part of the {\tsc{Pythia}} resonance machinery or as part of the {\tsc{Jetset}} particle decay one. The $\nu_{\chi}$ is assumed stable, so the option above would currently make no difference. \begin{subentry} \iteme{= 0 :} hadrons are first produced, which subsequently decay (in \ttt{LUDECY}). The new quark may be allowed to shower, see \ttt{MSTJ(27)}. The $\mathrm{W}$ produced in the decay is code $\pm89$. \iteme{= 1 :} the heavy quark first decays (in \ttt{PYRESD}) to a light one, and thereafter the light quark fragments. Parton showering in the decay is automatically included, but can be switched off with \ttt{MSTP(71)}. The $\mathrm{W}$ has the ordinary code $\pm24$, and is allowed to decay isotropically. \iteme{= 2 :} as \ttt{=1}, except that the $\mathrm{W}$ decay is anisotropic, as expected from $\mathrm{W}$ polarization in the heavy flavour decay. \itemc{Note:} in options 1 and 2 the cross section is reduced to take into account restrictions on allowed decay modes, while no such reduction occurs for option 0. See further section \ref{sss:heavflavclass}. \end{subentry} \iteme{MSTP(51) :} (D=9) choice of proton parton-distribution set; see also \ttt{MSTP(52)}. \begin{subentry} \iteme{= 1 :} EHLQ set 1 (1986 updated version). \iteme{= 2 :} EHLQ set 2 (1986 updated version). \iteme{= 3 :} Duke--Owens set 1. \iteme{= 4 :} Duke--Owens set 2. \iteme{= 5 :} CTEQ2M (best $\br{\mrm{MS}}$ fit). \iteme{= 6 :} CTEQ2MS (singular at small $x$). \iteme{= 7 :} CTEQ2MF (flat at small $x$). \iteme{= 8 :} CTEQ2ML (large $\Lambda$). \iteme{= 9 :} CTEQ2L (best leading order fit). \iteme{= 10 :} CTEQ2D (best DIS fit). \iteme{= 11 :} GRV LO (1992 updated version). \itemc{Note:} since all parametrizations have some region of applicability, the parton distributions are assumed frozen below the lowest $Q^2$ covered by the parametrizations; the CTEQ2 ones have been allowed to extend down to $Q_{\mathrm{min}} = 1$~GeV. For the former four, evolution is also frozed above the maximum $Q^2$. The extrapolation of EHLQ to low $x$ is covered by \ttt{PARP(51)}. \end{subentry} \iteme{MSTP(52) :} (D=1) choice of proton parton-distribution-function library. \begin{subentry} \iteme{= 1 :} the internal {\tsc{Pythia}} one, with parton distributions according to the \ttt{MSTP(51)} above. \iteme{= 2 :} the \tsc{Pdflib} one \cite{Plo93}, with the \tsc{Pdflib} (version 4) \ttt{NGROUP} and \ttt{NSET} numbers to be given as \ttt{MSTP(51) = 1000}$\times$\ttt{NGROUP + NSET}. \itemc{Note:} to make use of option 2, it is necessary to link \tsc{Pdflib}. Additionally, on most computers, the two dummy routines \ttt{PDFSET} and \ttt{STRUCTM} at the end of the {\tsc{Pythia}} file should be removed or commented out. \itemc{Warning:} For external parton distribution libraries, {\tsc{Pythia}} does not check whether \ttt{MSTP(51)} corresponds to a valid code, or if special $x$ and $Q^2$ restrictions exist for a given set, such that crazy values could be returned. This puts an extra responsibility on you. \end{subentry} \iteme{MSTP(53) :} (D=1) choice of pion parton-distribution set; see also \ttt{MSTP(54)}. \begin{subentry} \iteme{= 1 :} Owens set 1. \iteme{= 2 :} Owens set 2. \iteme{= 3 :} GRV LO (updated version). \end{subentry} \iteme{MSTP(54) :} (D=1) choice of pion parton-distribution-function library. \begin{subentry} \iteme{= 1 :} the internal {\tsc{Pythia}} one, with parton distributions according to the \ttt{MSTP(53)} above. \iteme{= 2 :} the \tsc{Pdflib} one \cite{Plo93}, with the \tsc{Pdflib} (version 4) \ttt{NGROUP} and \ttt{NSET} numbers to be given as \ttt{MSTP(53) = 1000}$\times$\ttt{NGROUP + NSET}. \itemc{Note:} to make use of option 2, it is necessary to link \tsc{Pdflib}. Additionally, on most computers, the two dummy routines \ttt{PDFSET} and \ttt{STRUCTM} at the end of the {\tsc{Pythia}} file should be removed or commented out. \itemc{Warning:} For external parton distribution libraries, {\tsc{Pythia}} does not check whether \ttt{MSTP(53)} corresponds to a valid code, or if special $x$ and $Q^2$ restrictions exist for a given set, such that crazy values could be returned. This puts an extra responsibility on you. \end{subentry} \iteme{MSTP(55)} : (D=5) choice of the parton-distribution set of the photon; see also \ttt{MSTP(56)}. \begin{subentry} \iteme{= 1 :} Drees--Grassie. \iteme{= 5 :} SaS 1D (in DIS scheme, with $Q_0=0.6$~GeV). \iteme{= 6 :} SaS 1M (in {$\overline{\mrm{MS}}$} scheme, with $Q_0=0.6$~GeV). \iteme{= 7 :} SaS 2D (in DIS scheme, with $Q_0=2$~GeV). \iteme{= 8 :} SaS 2M (in {$\overline{\mrm{MS}}$} scheme, with $Q_0=2$~GeV). \iteme{= 9 :} SaS 1D (in DIS scheme, with $Q_0=0.6$~GeV). \iteme{= 10 :} SaS 1M (in {$\overline{\mrm{MS}}$} scheme, with $Q_0=0.6$~GeV). \iteme{= 11 :} SaS 2D (in DIS scheme, with $Q_0=2$~GeV). \iteme{= 12 :} SaS 2M (in {$\overline{\mrm{MS}}$} scheme, with $Q_0=2$~GeV). \itemc{Note 1:} sets 5--8 use the parton distributions of the respective set, and nothing else. These are appropriate for most applications, e.g. jet production in $\gamma\mathrm{p}$ and $\gamma\gamma$ collisions. Sets 9--12 instead are appropriate for $\gamma^*\gamma$ processes, i.e. DIS scattering on a photon, as measured in $F_2^{\gamma}$. Here the anomalous contribution for $\c$ and $\b$ quarks are handled by the Bethe-Heitler formulae, and the direct term is artificially lumped with the anomalous one, so that the event simulation more closely agrees with what will be experimentally observed in these processes. The agreement with the $F_2^{\gamma}$ parametrization is still not perfect, e.g. in the treatment of heavy flavours close to threshold. \itemc{Note 2:} Sets 5--12 contain both VMD pieces and anomalous pieces, separately parametrized. Therefore the respective piece is automatically called, whatever \ttt{MSTP(14)} value is used to select only a part of the allowed photon interactions. For other sets (set 1 above or \tsc{Pdflib} sets), usually there is no corresponding subdivision. Then an option like \ttt{MSTP(14)=2} (VMD part of photon only) is based on a rescaling of the pion distributions, while \ttt{MSTP(14)=3} gives the SaS anomalous parametrization. \itemc{Note 3:} Formally speaking, the $k_0$ (or $p_0$) cut-off in \ttt{PARP(15)} need not be set in any relation to the $Q_0$ cut-off scales used by the various parametrizations. Indeed, due to the familiar scale choice ambiguity problem, there could well be some offset between the two. However, unless you know what you are doing, it is strongly recommended that you let the two agree, i.e. set \ttt{PARP(15)=0.6} for the SaS 1 sets and \ttt{=2.} for the SaS 2 sets. \end{subentry} \iteme{MSTP(56) :} (D=1) choice of photon parton-distribution-function library. \begin{subentry} \iteme{= 1 :} the internal {\tsc{Pythia}} one, with parton distributions according to the \ttt{MSTP(55)} above. \iteme{= 2 :} the \tsc{Pdflib} one \cite{Plo93}, with the \tsc{Pdflib} (version 4) \ttt{NGROUP} and \ttt{NSET} numbers to be given as \ttt{MSTP(55) = 1000}$\times$\ttt{NGROUP + NSET}. When the VMD and anomalous parts of the photon are split, like for \ttt{MSTP(14)=10}, it is necessary to specify pion set to be used for the VMD component, in \ttt{MSTP(53)} and \ttt{MSTP(54)}, while \ttt{MSTP(55)} here is irrelevant. \iteme{= 3 :} when the parton distributions of the anomalous photon are requested, the homogeneous solution is provided, evolved from a starting value \ttt{PARP(15)} to the requested $Q$ scale. The homogeneous solution is normalized so that the net momentum is unity, i.e. any factors of $\alpha_{\mathrm{em}}/2\pi$ and charge have been left out. The flavour of the original $\mathrm{q}$ is given in \ttt{MSTP(55)} (1, 2, 3, 4 or 5 for $\d$, $\u$, $\mathrm{s}$, $\c$ or $\b$); the value 0 gives a mixture according to squared charge, with the exception that $\c$ and $\b$ are only allowed above the respective mass threshold ($Q > m_{\mathrm{q}}$). The four-flavour $\Lambda$ value is assumed given in \ttt{PARP(1)}; it is automatically recalculated for 3 or 5 flavours at thresholds. This option is not intended for standard event generation, but is useful for some theoretical studies. \itemc{Note:} to make use of option 2, it is necessary to link \tsc{Pdflib}. Additionally, on most computers, the two dummy routines \ttt{PDFSET} and \ttt{STRUCTM} at the end of the {\tsc{Pythia}} file should be removed or commented out. \itemc{Warning:} For external parton-distribution libraries, {\tsc{Pythia}} does not check whether \ttt{MSTP(55)} corresponds to a valid code, or if special $x$ and $Q^2$ restrictions exist for a given set, such that crazy values could be returned. This puts an extra responsibility on you. \end{subentry} \iteme{MSTP(57) :} (D=1) choice of $Q^2$ dependence in parton-distribution functions. \begin{subentry} \iteme{= 0 :} parton distributions are evaluated at nominal lower cut-off value $Q_0^2$, i.e. are made $Q^2$-independent. \iteme{= 1 :} the parametrized $Q^2$ dependence is used. \iteme{= 2 :} the parametrized parton-distribution behaviour is kept at large $Q^2$ and $x$, but modified at small $Q^2$ and/or $x$, so that parton distributions vanish in the limit $Q^2 \to 0$ and have a theoretically motivated small-$x$ shape \cite{Sch93a}. This option is only valid for the $\mathrm{p}$ and $\mathrm{n}$. \iteme{= 3 :} as \ttt{=2}, except that also the $\pi^{\pm}$ and the VMD component of a photon is modified in a corresponding manner. \end{subentry} \iteme{MSTP(58) :} (D=min(6, 2$\times$\ttt{MSTP(1)})) maximum number of quark flavours used in parton distributions, and thus also for initial-state space-like showers. If some distributions (notably $\t$) are absent in the parametrization selected in \ttt{MSTP(51)}, these are obviously automatically excluded. \iteme{MSTP(61) :} (D=1) (C) master switch for initial-state QCD and QED radiation. \begin{subentry} \iteme{= 0 :} off. \iteme{= 1 :} on. \end{subentry} \iteme{MSTP(62) - MSTP(67) :} (C) further switches for initial-state radiation, see section \ref{ss:showrout}. \iteme{MSTP(71) :} (D=1) (C) master switch for final-state QCD and QED radiation. \begin{subentry} \iteme{= 0 :} off. \iteme{= 1 :} on. \itemc{Note:} additional switches (e.g. for conventional/coherent showers) are available in \ttt{MSTJ(40) - MSTJ(50)} and \ttt{PARJ(81) - PARJ(89)}, see section \ref{ss:showrout}. \end{subentry} \iteme{MSTP(81) :} (D=1) master switch for multiple interactions. \begin{subentry} \iteme{= 0 :} off. \iteme{= 1 :} on. \end{subentry} \iteme{MSTP(82) - MSTP(83) :} further switches for multiple interactions, see section \ref{ss:multintpar}. \iteme{MSTP(85) :} possibility to dampen hard scattering matrix elements in the limit $p_{\perp} \to 0$. It parellels some of the multiple interactions options for QCD processes, but can be used for any $2 \to 2$ process. \begin{subentry} \iteme{= 0 :} off, i.e. standard matrix elements are kept. \iteme{= 1 :} on, i.e. matrix elements are multiplied by a factor $p_{\perp}^4/(p_{\perp}^2 + p_{\perp 0}^2)^2$, where $p_{\perp 0}$ is given by \ttt{PARP(82)}. Additionally $\alpha_{\mathrm{s}}$ is evaluated at a scale $p_{\perp}^2 + p_{\perp 0}^2$ rather than just $p_{\perp}^2$. \end{subentry} \iteme{MSTP(91) - MSTP(94) :} switches for beam remnant treatment, see section \ref{ss:multintpar}. \iteme{MSTP(101) :} (D=3) (C) structure of diffractive system. \begin{subentry} \iteme{= 1 :} forward moving diquark + interacting quark. \iteme{= 2 :} forward moving diquark + quark joined via interacting gluon (`hairpin' configuration). \iteme{= 3 :} a mixture of the two options above, with a fraction \ttt{PARP(101)} of the former type. \end{subentry} \iteme{MSTP(102) :} (D=1) (C) decay of a $\rho^0$ meson produced by `elastic' scattering of an incoming $\gamma$, as in $\gamma \mathrm{p} \to \rho^0 \mathrm{p}$, or the same with the hadron diffractively excited. \begin{subentry} \iteme{= 0 :} the $\rho^0$ is allowed to decay isotropically, like any other $\rho^0$. \iteme{= 1 :} the decay $\rho^0 \to \pi^+ \pi^-$ is done with an angular distribution proportional to $\sin^2 \theta$ in its rest frame, where the $z$ axis is given by the direction of motion of the $\rho^0$. The $\rho^0$ decay is then done as part of the hard process, i.e. also when \ttt{MSTP(111)=0}. \end{subentry} \iteme{MSTP(111) :} (D=1) (C) master switch for fragmentation and decay, as obtained with a \ttt{LUEXEC} call. \begin{subentry} \iteme{= 0 :} off. \iteme{= 1 :} on. \iteme{= -1 :} only choose kinematical variables for hard scattering, i.e. no jets are defined. This is useful, for instance, to calculate cross sections (by Monte Carlo integration) without wanting to simulate events; information obtained with \ttt{PYSTAT(1)} will be correct. \end{subentry} \iteme{MSTP(112) :} (D=1) (C) cuts on partonic events; only affects an exceedingly tiny fraction of events. \begin{subentry} \iteme{= 0 :} no cuts (can be used only with independent fragmentation, at least in principle). \iteme{= 1 :} string cuts (as normally required for fragmentation). \end{subentry} \iteme{MSTP(113) :} (D=1) (C) recalculation of energies of partons from their momenta and masses, to be done immediately before and after fragmentation, to partly compensate for some numerical problems appearing at high energies. \begin{subentry} \iteme{= 0 :} not performed. \iteme{= 1 :} performed. \end{subentry} \iteme{MSTP(121) :} (D=0) calculation of kinematics selection coefficients and differential cross section maxima for included (by user or default) subprocesses. \begin{subentry} \iteme{= 0 :} not known; to be calculated at initialization. \iteme{= 1 :} not known; to be calculated at initialization; however, the maximum value then obtained is to be multiplied by \ttt{PARP(121)} (this may be useful if a violation factor has been observed in a previous run of the same kind). \iteme{= 2 :} known; kinematics selection coefficients stored by user in \ttt{COEF(ISUB,J)} (\ttt{J} = 1--20) in common block \ttt{PYINT2} and maximum of the corresponding differential cross section times Jacobians in \ttt{XSEC(ISUB,1)} in common block \ttt{PYINT5}. This is to be done for each included subprocess ISUB before initialization, with the sum of all \ttt{XSEC(ISUB,1)} values, except for ISUB = 95, stored in \ttt{XSEC(0,1)}. \end{subentry} \iteme{MSTP(122) :} (D=1) initialization and differential cross section maximization print-out. \begin{subentry} \iteme{= 0 :} none. \iteme{= 1 :} short message. \iteme{= 2 :} detailed message, including full maximization. \end{subentry} \iteme{MSTP(123) :} (D=2) reaction to violation of maximum differential cross section. \begin{subentry} \iteme{= 0 :} stop generation, print message. \iteme{= 1 :} continue generation, print message for each subsequently larger violation. \iteme{= 2 :} as \ttt{=1}, but also increase value of maximum. \end{subentry} \iteme{MSTP(124) :} (D=1) (C) frame for presentation of event. \begin{subentry} \iteme{= 1 :} as specified in \ttt{PYINIT}. \iteme{= 2 :} c.m. frame of incoming particles. \end{subentry} \iteme{MSTP(125) :} (D=1) (C) documentation of partonic process, see section \ref{sss:PYrecord} for details. \begin{subentry} \iteme{= 0 :} only list ultimate string/particle configuration. \iteme{= 1 :} additionally list short summary of the hard process. \iteme{= 2 :} list complete documentation of intermediate steps of parton-shower evolution. \end{subentry} \iteme{MSTP(126) :} (D=20) number of lines at the beginning of event record that are reserved for event-history information; see section \ref{sss:PYrecord}. This value should never be reduced, but may be increased at a later date if more complicated processes are included. \iteme{MSTP(128) :} (D=0) storing of copy of resonance decay products in the documentation section of the event record, and mother pointer (\ttt{K(I,3)}) relation of the actual resonance decay products (stored in the main section of the event record) to the documentation copy. \begin{subentry} \iteme{= 0 :} products are stored also in the documentation section, and each product stored in the main section points back to the corresponding entry in the documentation section. \iteme{= 1 :} products are stored also in the documentation section, but the products stored in the main section point back to the decaying resonance copy in the main section. \iteme{= 2 :} products are not stored in the documentation section; the products stored in the main section point back to the decaying resonance copy in the main section. \end{subentry} \iteme{MSTP(129) :} (D=10) for the maximization of $2 \to 3$ processes (\ttt{ISET(ISUB)=5}) each phase-space point in $\tau$, $y$ and $\tau'$ is tested \ttt{MSTP(129)} times in the other dimensions (at randomly selected points) to determine the effective maximum in the ($\tau$, $y$, $\tau'$) point. \iteme{MSTP(131) :} (D=0) master switch for pile-up events, i.e. several independent hadron--hadron interactions generated in the same bunch--bunch crossing, with the events following one after the other in the event record. \begin{subentry} \iteme{= 0 :} off, i.e. only one event is generated at a time. \iteme{= 1 :} on, i.e. several events are allowed in the same event record. Information on the processes generated may be found in \ttt{MSTI(41) - MSTI(50)}. \end{subentry} \iteme{MSTP(132) - MSTP(134) :} further switches for pile-up events, see section \ref{ss:multintpar}. \iteme{MSTP(141) :} (D=0) calling of \ttt{PYKCUT} in the event-generation chain, for inclusion of user-specified cuts. \begin{subentry} \iteme{= 0 :} not called. \iteme{= 1 :} called. \end{subentry} \iteme{MSTP(142) :} (D=0) calling of \ttt{PYEVWT} in the event-generation chain, either to give weighted events or to modify standard cross sections. See \ttt{PYEVWT} description in section \ref{ss:PYTmainroutines} for further details. \begin{subentry} \iteme{= 0 :} not called. \iteme{= 1 :} called; the distribution of events among subprocesses and in kinematics variables is modified by the factor \ttt{WTXS}, set by the user in the \ttt{PYEVWT} call, but events come with a compensating weight \ttt{PARI(10)=1./WTXS}, such that total cross sections are unchanged. \iteme{= 2 :} called; the cross section itself is modified by the factor \ttt{WTXS}, set by the user in the \ttt{PYEVWT} call. \end{subentry} \iteme{MSTP(151) :} (D=0) introduce smeared position of primary vertex of events. \begin{subentry} \iteme{= 0 :} no, i.e. the primary vertex of each event is at the origin. \iteme{= 1 :} yes, with Gaussian distributions separately in $x$, $y$, $z$ and $t$. The respective widths of the Gaussians have to be given in \ttt{PARP(151) - PARP(154)}. Also pile-up events obtain separate primary vertices. No provisions are made for more complicated beam-spot shapes, e.g. with a spread in $z$ that varies as a function of $t$. Note that a large beam spot combined with some of the \ttt{MSTJ(22)} options may lead to many particles not being allowed to decay at all. \end{subentry} \iteme{MSTP(171) :} (D=0) possibility of variable energies from one event to the next. For further details see subsection \ref{ss:PYvaren}. \begin{subentry} \iteme{= 0 :} no; i.e. the energy is fixed at the initialization call. \iteme{= 1 :} yes; i.e. a new energy has to be given for each new event. \end{subentry} \iteme{MSTP(172) :} (D=2) options for generation of events with variable energies, applicable when \ttt{MSTP(171)=1}. \begin{subentry} \iteme{= 1 :} an event is generated at the requested energy, i.e. internally a loop is performed over possible event configurations until one is accepted. If the requested c.m. energy of an event is below \ttt{PARP(2)} the run is aborted. Cross-section information can not be trusted with this option, since it depends on how you decided to pick the requested energies. \iteme{= 2 :} only one event configuration is tried. If that is accepted, the event is generated in full. If not, no event is generated, and the status code \ttt{MSTI(61)=1} is returned. You are then expected to give a new energy, looping until an acceptable event is found. No event is generated if the requested c.m. energy is below \ttt{PARP(2)}, instead \ttt{MSTI(61)=1} is set to signal the failure. In principle, cross sections should come out correctly with this option. \end{subentry} \iteme{MSTP(173) :} (D=0) possibility for user to give in an event weight to compensate for a biased choice of beam spectrum. \begin{subentry} \iteme{= 0 :} no, i.e. event weight is unity. \iteme{= 1 :} yes; weight to be given for each event in \ttt{PARP(173)}, with maximum weight given at initialization in \ttt{PARP(174)}. \end{subentry} \iteme{MSTP(181) :} (R) {\tsc{Pythia}} version number. \iteme{MSTP(182) :} (R) {\tsc{Pythia}} subversion number. \iteme{MSTP(183) :} (R) last year of change for {\tsc{Pythia}}. \iteme{MSTP(184) :} (R) last month of change for {\tsc{Pythia}}. \iteme{MSTP(185) :} (R) last day of change for {\tsc{Pythia}}. \iteme{MSTP(186) :} (R) earliest subversion of {\tsc{Jetset}} version 7 with which this {\tsc{Pythia}} subversion can be run. \boxsep \iteme{PARP(1) :}\label{p:PARP} (D=0.25 GeV) nominal $\Lambda_{\mrm{QCD}}$ used in running $\alpha_{\mathrm{s}}$ for hard scattering (see \ttt{MSTP(3)}). \iteme{PARP(2) :} (D=10. GeV) lowest c.m. energy for the event as a whole that the program will accept to simulate. \iteme{PARP(13) :} (D=1. GeV$^2$) $Q_{\mathrm{max}}^2$ scale, to be set by user for defining maximum scale allowed for photoproduction when using the option \ttt{MSTP(13)=2}. \iteme{PARP(14) :} (D=0.01) in the numerical integration of quark and gluon parton distributions inside an electron, the successive halvings of evaluation-point spacing is interrupted when two values agree in relative size, $|$new$-$old$|$/(new$+$old), to better than \ttt{PARP(14)}. There are hardwired lower and upper limits of 2 and 8 halvings, respectively. \iteme{PARP(15) :} (D=0.5 GeV) lower cut-off $p_0$ used to define minimum transverse momentum in branchings $\gamma \to \mathrm{q}\overline{\mathrm{q}}$ in the anomalous event class of $\gamma\mathrm{p}$ interactions. \iteme{PARP(16) :} (D=1.) the anomalous parton-distribution functions of the photon are taken to have the charm and bottom flavour thresholds at virtuality \ttt{PARP(16)}$\times m_{\mathrm{q}}^2$. \iteme{PARP(17) :} (D=1.) rescaling factor used for the $Q$ argument of the anomalous parton distributions of the photon, see \ttt{MSTP(15)}. \iteme{PARP(31) :} (D=1.5) common $K$ factor multiplying the differential cross section for hard parton--parton processes when \ttt{MSTP(33)=1} or \ttt{2}, with the exception of colour annihilation graphs in the latter case. \iteme{PARP(32) :} (D=2.0) special $K$ factor multiplying the differential cross section in hard colour annihilation graphs, including resonance production, when \ttt{MSTP(33)=2}. \iteme{PARP(33) :} (D=0.075) this factor is used to multiply the ordinary $Q^2$ scale in $\alpha_{\mathrm{s}}$ at the hard interaction for \ttt{MSTP(33)=3}. The effective $K$ factor thus obtained is in accordance with the results in \cite{Ell86}. \iteme{PARP(34) :} (D=1.) the $Q^2$ scale defined by \ttt{MSTP(32)} is multiplied by \ttt{PARP(34)} when it is used as argument for parton distributions and $\alpha_{\mathrm{s}}$ at the hard interaction. It does not affect $\alpha_{\mathrm{s}}$ when \ttt{MSTP(33)=3}, nor does it change the $Q^2$ argument of parton showers. \iteme{PARP(35) :} (D=0.20) fix $\alpha_{\mathrm{s}}$ value that is used in the heavy-flavour threshold factor when \ttt{MSTP(35)=1}. \iteme{PARP(36) :} (D=0. GeV) the width $\Gamma_{\mathrm{Q}}$ for the heavy flavour studied in processes ISUB = 81 or 82; to be used for the threshold factor when \ttt{MSTP(35)=2}. \iteme{PARP(37) :} (D=2.) for \ttt{MSTP(37)=1} this regulates the point at which the reference on-shell quark mass in Higgs couplings is assumed defined; specifically the running quark mass is assumed to coincide with the fix one at an energy scale \ttt{PARP(37)} times the fix quark mass, i.e. $m_{\mrm{running}}($\ttt{PARP(37)}$\times m_{\mrm{fix}}) = m_{\mrm{fix}}$. \iteme{PARP(38) :} (D=0.70 GeV$^3$) the squared wave function at the origin, $|R(0)|^2$, of the $\mathrm{J}/\psi$ wave function. Used for process 86. See ref. \cite{Glo88}. \iteme{PARP(39) :} (D=0.006 GeV$^3$) the squared derivative of the wave function at the origin, $|R'(0)|^2/m^2$, of the $\chi_{\c}$ wave functions. Used for the processes 87, 88 and 89. See ref. \cite{Glo88}. \iteme{PARP(41) :} (D=0.020 GeV) in the process of generating mass for resonances, and optionally to force that mass to be in a given range, only resonances with a total width in excess of \ttt{PARP(41)} are generated according to a Breit--Wigner shape (if allowed by \ttt{MSTP(42)}), while narrower resonances are put on the mass shell. \iteme{PARP(42) :} (D=2. GeV) minimum mass of resonances assumed to be allowed when evaluating total width of $\H^0$ to $\mathrm{Z}^0 \mathrm{Z}^0$ or $\mathrm{W}^+ \mathrm{W}^-$ for cases when the $\H^0$ is so light that (at least) one $\mathrm{Z}/\mathrm{W}$ is forced to be off the mass shell. Also generally used as safety check on minimum mass of resonance. Note that some \ttt{CKIN} values may provide additional constraints. \iteme{PARP(43) :} (D=0.10) precision parameter used in numerical integration of width into channel with at least one daughter off the mass shell. \iteme{PARP(44) :} (D=1000.) the $\nu$ parameter of the strongly interacting $\mathrm{Z}/\mathrm{W}$ model of Dobado, Herrero and Terron \cite{Dob91}. \iteme{PARP(45) :} (D=2054. GeV) the effective techni-$\rho$ mass parameter of the strongly interacting model of Dobado, Herrero and Terron \cite{Dob91}; see \ttt{MSTP(46)=5}. On physical grounds it should not be chosen smaller than about 1 TeV or larger than about the default value. \iteme{PARP(46) :} (D=123. GeV) the $F_{\pi}$ decay constant that appears inversely quadratically in all techni-$\eta$ partial decay widths \cite{Eic84,App92}. \iteme{PARP(47) :} (D=246. GeV) vacuum expectation value $v$ used in the DHT scenario \cite{Dob91} to define the width of the techni-$\rho$; this width is inversely proportional $v^2$. \iteme{PARP(51) :} (D=1.) if parton distributions for light flavours have to be extrapolated to $x$ values lower than covered by the parametrizations, an $x^{-b}$ behaviour, with $b =$\ttt{PARP(51)}, is assumed in that region. This option only applies for the EHLQ proton parton distributions that are internal to {\tsc{Pythia}}. \iteme{PARP(61) - PARP(65) :} (C) parameters for initial-state radiation, see section \ref{ss:showrout}. \iteme{PARP(71) - PARP(72) :} (C) parameter for final-state radiation, see section \ref{ss:showrout}. \iteme{PARP(81) - PARP(88) :} parameters for multiple interactions, see section \ref{ss:multintpar}. \iteme{PARP(91) - PARP(100) :} parameters for beam remnant treatment, see section \ref{ss:multintpar}. \iteme{PARP(101) :} (D=0.50) fraction of diffractive systems in which a quark is assumed kicked out by the pomeron rather than a gluon; applicable for option \ttt{MSTP(101)=3}. \iteme{PARP(102) :} (D=0.28 GeV) the mass spectrum of diffractive states (in single and double diffractive scattering) is assumed to start \ttt{PARP(102)} above the mass of the particle that is diffractively excited. In this connection, an incoming $\gamma$ is taken to have the selected VMD meson mass, i.e. $m_{\rho}$, $m_{\omega}$, $m_{\phi}$ or $m_{\mathrm{J}/\psi}$. \iteme{PARP(103) :} (D=1.0 GeV) if the mass of a diffractive state is less than \ttt{PARP(103)} above the mass of the particle that is diffractively excited, the state is forced to decay isotropically into a two-body channel. In this connection, an incoming $\gamma$ is taken to have the selected VMD meson mass, i.e. $m_{\rho}$, $m_{\omega}$, $m_{\phi}$ or $m_{\mathrm{J}/\psi}$. If the mass is higher than this threshold, the standard string fragmentation machinery is used. The forced two-body decay is always carried out, also when \ttt{MSTP(111)=0}. \iteme{PARP(111) :} (D=2. GeV) used to define the minimum invariant mass of the remnant hadronic system (i.e. when interacting partons have been taken away), together with original hadron masses and extra parton masses. \iteme{PARP(121) :} (D=1.) the maxima obtained at initial maximization are multiplied by this factor if \ttt{MSTP(121)=1}; typically \ttt{PARP(121)} would be given as the product of the violation factors observed (i.e. the ratio of final maximum value to initial maximum value) for the given process(es). \iteme{PARP(122) :} (D=0.4) fraction of total probability that is shared democratically between the \ttt{COEF} coefficients open for the given variable, with the remaining fraction distributed according to the optimization results of \ttt{PYMAXI}. \iteme{PARP(131) :} parameter for pile-up events, see section \ref{ss:multintpar}. \iteme{PARP(151) - PARP(154) :} (D=4*0.) (C) regulate the assumed beam-spot size. For \ttt{MSTP(151)=1} the $x$, $y$, $z$ and $t$ coordinates of the primary vertex of each event are selected according to four independent Gaussians. The widths of these Gaussians are given by the four parameters, where the first three are in units of mm and the fourth in mm/$c$. \iteme{PARP(161) - PARP(164) :} (D=2.20, 23.6, 18.4, 11.5) couplings $f_V^2/4\pi$ of the photon to the $\rho^0$, $\omega$, $\phi$ and $\mathrm{J}/\psi$ vector mesons. \iteme{PARP(171) :} to be set, event-by-event, when variable energies are allowed, i.e. when \ttt{MSTP(171)=1}. If \ttt{PYINIT} is called with \ttt{FRAME='CMS'} (\ttt{='FIXT'}), \ttt{PARP(171)} multiplies the c.m. energy (beam energy) used at initialization. For the options \ttt{'USER'}, \ttt{'FOUR'} and \ttt{'FIVE'}, \ttt{PARP(171)} is dummy, since there the momenta are set in the \ttt{P} array. \iteme{PARP(173) :} event weight to be given by user when \ttt{MSTP(173)=1}. \iteme{PARP(174) :} (D=1.) maximum event weight that will be encountered in \ttt{PARP(173)} during the course of a run with \ttt{MSTP(173)=1}; to be used to optimize the efficiency of the event generation. It is always allowed to use a larger bound than the true one, but with a corresponding loss in efficiency. \end{entry} \subsection{General Event Information} When an event is generated with \ttt{PYEVNT}, some information on it is stored in the \ttt{MSTI} and \ttt{PARI} arrays of the \ttt{PYPARS} common block (often copied directly from the internal \ttt{MINT} and \ttt{VINT} variables). Further information is stored in the complete event record; see section \ref{ss:evrec}. Part of the information is only relevant for some subprocesses; by default everything irrelevant is set to 0. Kindly note that, like the \ttt{CKIN} constraints described in section \ref{ss:PYswitchkin}, kinematical variables normally (i.e. where it is not explicitly stated otherwise) refer to the na\"{\i}ve hard scattering, before initial- and final-state radiation effects have been included. \drawbox{COMMON/PYPARS/MSTP(200),PARP(200),MSTI(200),PARI(200)}% \label{p:PYPARS2} \begin{entry} \itemc{Purpose:} to provide information on latest event generated or, in a few cases, on statistics accumulated during the run. \iteme{MSTI(1) :}\label{p:MSTI} specifies the general type of subprocess that has occurred, according to the ISUB code given in section \ref{ss:ISUBcode}. \iteme{MSTI(2) :} whenever \ttt{MSTI(1)} (together with \ttt{MSTI(15)} and \ttt{MSTI(16)}) are not enough to specify the type of process uniquely, \ttt{MSTI(2)} provides an ordering of the different possibilities. This is particularly relevant for the different colour-flow topologies possible in QCD $2 \to 2$ processes. With $i = $\ttt{MSTI(15)}, $j = $\ttt{MSTI(16)} and $k = $\ttt{MSTI(2)}, the QCD possibilities are, in the classification scheme of \cite{Ben84} (cf. section \ref{sss:QCDjetclass}): \begin{subentry} \itemn{ISUB = 11,} $i = j$, $\mathrm{q}_i \mathrm{q}_i \to \mathrm{q}_i \mathrm{q}_i$; \\ $k = 1$ : colour configuration $A$. \\ $k = 2$ : colour configuration $B$. \itemn{ISUB = 11,} $i \neq j$, $\mathrm{q}_i \mathrm{q}_j \to \mathrm{q}_i \mathrm{q}_j$; \\ $k = 1$ : only possibility. \itemn{ISUB = 12,} $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{q}_l \overline{\mathrm{q}}_l$; \\ $k = 1$ : only possibility. \itemn{ISUB = 13,} $\mathrm{q}_i \overline{\mathrm{q}}_i \to \mathrm{g} \mathrm{g}$; \\ $k = 1$ : colour configuration $A$. \\ $k = 2$ : colour configuration $B$. \itemn{ISUB = 28,} $\mathrm{q}_i \mathrm{g} \to \mathrm{q}_i \mathrm{g}$; \\ $k = 1$ : colour configuration $A$. \\ $k = 2$ : colour configuration $B$. \itemn{ISUB = 53,} $\mathrm{g} \mathrm{g} \to \mathrm{q}_l \overline{\mathrm{q}}_l$; \\ $k = 1$ : colour configuration $A$. \\ $k = 2$ : colour configuration $B$. \itemn{ISUB = 68,} $\mathrm{g} \mathrm{g} \to \mathrm{g} \mathrm{g}$; \\ $k = 1$ : colour configuration $A$. \\ $k = 2$ : colour configuration $B$. \\ $k = 3$ : colour configuration $C$. \itemn{ISUB = 83,} $\mathrm{f} \mathrm{q} \to \mathrm{f}' \mathrm{Q}$ (by $t$-channel $\mathrm{W}$ exchange; does not distinguish colour flows but result of user selection); \\ $k = 1$ : heavy flavour $\mathrm{Q}$ is produced on side 1. \\ $k = 2$ : heavy flavour $\mathrm{Q}$ is produced on side 2. \end{subentry} \iteme{MSTI(3) :} the number of partons produced in the hard interactions, i.e. the number $n$ of the $2 \to n$ matrix elements used; it is sometimes 3 or 4 when a basic $2 \to 1$ or $2 \to 2$ process has been folded with two $1 \to 2$ initial branchings (like $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \mathrm{q}''' \H^0$). \iteme{MSTI(4) :} number of documentation lines at the beginning of the common block \ttt{LUJETS} that are given with \ttt{K(I,1)=21}; 0 for \ttt{MSTP(125)=0}. \iteme{MSTI(5) :} number of events generated to date in current run. In runs with the variable-energy option, \ttt{MSTP(171)=1} and \ttt{MSTP(172)=2}, only those events that survive (i.e. that do not have \ttt{MSTI(61)=1}) are counted in this number. That is, \ttt{MSTI(5)} may be less than the total number of \ttt{PYEVNT} calls. \iteme{MSTI(6) :} current frame of event, cf. \ttt{MSTP(124)}. \iteme{MSTI(7), MSTI(8) :} line number for documentation of outgoing partons/particles from hard scattering for $2 \to 2$ or $2 \to 1 \to 2$ processes (else = 0). \iteme{MSTI(9) :} event class used in current event for $\gamma\mathrm{p}$ or $\gamma\gamma$ events generated with the \ttt{MSTP(14)=10} option. \begin{subentry} \iteme{= 0 :} for other processes than the ones listed above. \iteme{= 1 :} VMD (for $\gamma\mathrm{p}$) or VMD*VMD (for $\gamma\gamma$). \iteme{= 2 :} direct (for $\gamma\mathrm{p}$) or VMD*direct (for $\gamma\gamma$). \iteme{= 3 :} anomalous (for $\gamma\mathrm{p}$) or VMD*anomalous (for $\gamma\gamma$). \iteme{= 4 :} direct*direct (for $\gamma\gamma$). \iteme{= 5 :} direct*anomalous (for $\gamma\gamma$). \iteme{= 6 :} anomalous*anomalous (for $\gamma\gamma$). \end{subentry} \iteme{MSTI(10) :} is 1 if cross section maximum was violated in current event, and 0 if not. \iteme{MSTI(11) :} KF flavour code for beam (side 1) particle. \iteme{MSTI(12) :} KF flavour code for target (side 2) particle. \iteme{MSTI(13), MSTI(14) :} KF flavour codes for side 1 and side 2 initial-state shower initiators. \iteme{MSTI(15), MSTI(16) :} KF flavour codes for side 1 and side 2 incoming partons to the hard interaction. \iteme{MSTI(17), MSTI(18) :} flag to signal if particle on side 1 or side 2 has been scattered diffractively; 0 if no, 1 if yes. \iteme{MSTI(21) - MSTI(24) :} KF flavour codes for outgoing partons from the hard interaction. The number of positions actually used is process-dependent, see \ttt{MSTI(3)}; trailing positions not used are set = 0. \iteme{MSTI(25), MSTI(26) :} KF flavour codes of the products in the decay of a single $s$-channel resonance formed in the hard interaction. Are thus only used when \ttt{MSTI(3)=1} and the resonance is allowed to decay. \iteme{MSTI(31) :} number of hard or semi-hard scatterings that occurred in the current event in the multiple-interaction scenario; is = 0 for a low-$p_{\perp}$ event. \iteme{MSTI(41) :} the number of pile-up events generated in the latest \ttt{PYEVNT} call (including the first, `hard' event). \iteme{MSTI(42) - MSTI(50) :} ISUB codes for the events 2--10 generated in the pile-up-events scenario. The first event ISUB code is stored in \ttt{MSTI(1)}. If \ttt{MSTI(41)} is less than 10, only as many positions are filled as there are pile-up events. If MSTI(41) is above 10, some ISUB codes will not appear anywhere. \iteme{MSTI(51) :} normally 0 but set to 1 if a \ttt{PYUPEV} call did not return an event, such that \ttt{PYEVNT} could not generate an event. For further details, see end of section \ref{ss:PYnewproc}. \iteme{MSTI(52) :} counter for the number of times the current event configuration failed in the generation machinery. For accepted events this is always 0, but the counter can be used inside \ttt{PYUPEV} to check on anomalous occurrences. For further details, see end of section \ref{ss:PYnewproc}. \iteme{MSTI(61) :} status flag set when events are generated. It is only of interest for runs with variable energies, \ttt{MSTP(171)=1}, with the option \ttt{MSTP(172)=2}. \begin{subentry} \iteme{= 0 :} an event has been generated. \iteme{= 1 :} no event was generated, either because the c.m. energy was too low or because the Monte Carlo phase space point selection machinery rejected the trial point. A new energy is to be picked by the user. \end{subentry} \boxsep \iteme{PARI(1) :}\label{p:PARI} total integrated cross section for the processes under study, in mb. This number is obtained as a by-product of the selection of hard-process kinematics, and is thus known with better accuracy when more events have been generated. The value stored here is based on all events until the latest one generated. \iteme{PARI(2) :} is the ratio \ttt{PARI(1)/MSTI(5)}, i.e. the ratio of total integrated cross section and number of events generated. Histograms filled with unit event weight have to be multiplied by this factor, at the end of the run, to convert results to mb. For \ttt{MSTP(142)=1}, \ttt{MSTI(5)} is replaced by the sum of \ttt{PARI(10)} values. Histograms are then filled with weight \ttt{PARI(10)} for each event and multiplied by \ttt{PARI(2)} at the end. In runs with the variable-energy option, \ttt{MSTP(171)=1} and \ttt{MSTP(172)=2}, only those events that survive (i.e. that do not have \ttt{MSTI(61)=1}) are counted. calls. \iteme{PARI(9) :} is weight \ttt{WTXS} returned from \ttt{PYEVWT} call when \ttt{MSTP(142)}$\geq 1$, otherwise is 1. \iteme{PARI(10) :} is compensating weight \ttt{1./WTXS} that should be associated to events when \ttt{MSTP(142)=1}, else is 1. \iteme{PARI(11) :} $E_{\mrm{cm}}$, i.e. total c.m. energy. \iteme{PARI(12) :} $s$, i.e. squared total c.m. energy. \iteme{PARI(13) :} $\hat{m} = \sqrt{\hat{s}}$, i.e. mass of the hard-scattering subsystem. \iteme{PARI(14) :} $\hat{s}$ of the hard subprocess ($2 \to 2$ or $2 \to 1$). \iteme{PARI(15) :} $\hat{t}$ of the hard subprocess ($2 \to 2$ or $2 \to 1 \to 2$). \iteme{PARI(16) :} $\hat{u}$ of the hard subprocess ($2 \to 2$ or $2 \to 1 \to 2$). \iteme{PARI(17) :} $\hat{p}_{\perp}$ of the hard subprocess ($2 \to 2$ or $2 \to 1 \to 2$), evaluated in the rest frame of the hard interaction. \iteme{PARI(18) :} $\hat{p}_{\perp}^2$ of the hard subprocess; see \ttt{PARI(17)}. \iteme{PARI(19) :} $\hat{m}'$, the mass of the complete three- or four-body final state in $2 \to 3$ or $2 \to 4$ processes (while $\hat{m}$, given in \ttt{PARI(13)}, here corresponds to the one- or two-body central system). Kinematically $\hat{m} \leq \hat{m}' \leq E_{\mrm{cm}}$. \iteme{PARI(20) :} $\hat{s}' = \hat{m}'^2$; see \ttt{PARI(19)}. \iteme{PARI(21) :} $Q$ of the hard-scattering subprocess. The exact definition is process-dependent, see \ttt{MSTP(32)}. \iteme{PARI(22) :} $Q^2$ of the hard-scattering subprocess; see \ttt{PARI(21)}. \iteme{PARI(23) :} $Q$ of the outer hard-scattering subprocess. Agrees with \ttt{PARI(21)} for a $2 \to 1$ or $2 \to 2$ process. For a $2 \to 3$ or $2 \to 4$ $\mathrm{W}/\mathrm{Z}$ fusion process, it is set by the $\mathrm{W}/\mathrm{Z}$ mass scale, and for subprocesses 121 and 122 by the heavy-quark mass. \iteme{PARI(24) :} $Q^2$ of the outer hard-scattering subprocess; see \ttt{PARI(23)}. \iteme{PARI(25) :} $Q$ scale used as maximum virtuality in parton showers. Is equal to \ttt{PARI(23)}, except for deep-inelastic-scattering processes when \ttt{MSTP(22)}$\geq 1$. \iteme{PARI(26) :} $Q^2$ scale in parton showers; see \ttt{PARI(25)}. \iteme{PARI(31), PARI(32) :} the momentum fractions $x$ of the initial-state parton-shower initiators on side 1 and 2, respectively. \iteme{PARI(33), PARI(34) :} the momentum fractions $x$ taken by the partons at the hard interaction, as used e.g. in the parton-distribution functions. \iteme{PARI(35) :} Feynman-$x$, $x_{\mrm{F}} = x_1 - x_2 = $\ttt{PARI(33)}$-$\ttt{PARI(34)}. \iteme{PARI(36) :} $\tau = \hat{s}/s = x_1 \, x_2 = $\ttt{PARI(33)}$\times$\ttt{PARI(34)}. \iteme{PARI(37) :} $y = (1/2) \ln(x_1/x_2)$, i.e. rapidity of the hard-interaction subsystem in the c.m. frame of the event as a whole. \iteme{PARI(38) :} $\tau' = \hat{s}'/s = $\ttt{PARI(20)/PARI(12)}. \iteme{PARI(39), PARI(40) :} the primordial $k_{\perp}$ values selected in the two beam remnants. \iteme{PARI(41) :} $\cos\hat{\theta}$, where $\hat{\theta}$ is the scattering angle of a $2 \to 2$ (or $2 \to 1 \to 2$) interaction, defined in the rest frame of the hard-scattering subsystem. \iteme{PARI(42) :} $x_{\perp}$, i.e. scaled transverse momentum of the hard-scattering subprocess, $x_{\perp} = 2 \hat{p}_{\perp}/E_{\mrm{cm}}$. \iteme{PARI(43), PARI(44) :} $x_{L3}$ and $x_{L4}$, i.e. longitudinal momentum fractions of the two scattered partons, in the range $-1 < x_{\mrm{L}} < 1$, in the c.m. frame of the event as a whole. \iteme{PARI(45), PARI(46) :} $x_3$ and $x_4$, i.e. scaled energy fractions of the two scattered partons, in the c.m. frame of the event as a whole. \iteme{PARI(47), PARI(48) :} $y^*_3$ and $y^*_4$, i.e. rapidities of the two scattered partons in the c.m. frame of the event as a whole. \iteme{PARI(49), PARI(50) :} $\eta^*_3$ and $\eta^*_4$, i.e. pseudorapidities of the two scattered partons in the c.m. frame of the event as a whole. \iteme{PARI(51), PARI(52) :} $\cos\theta^*_3$ and $\cos\theta^*_4$, i.e. cosines of the polar angles of the two scattered partons in the c.m. frame of the event as a whole. \iteme{PARI(53), PARI(54) :} $\theta^*_3$ and $\theta^*_4$, i.e. polar angles of the two scattered partons, defined in the range $0 < \theta^* < \pi$, in the c.m. frame of the event as a whole. \iteme{PARI(55), PARI(56) :} azimuthal angles $\phi^*_3$ and $\phi^*_4$ of the two scattered partons, defined in the range $-\pi < \phi^* < \pi$, in the c.m. frame of the event as a whole. \iteme{PARI(61) :} multiple interaction enhancement factor for current event. A large value corresponds to a central collision and a small value to a peripheral one. \iteme{PARI(65) :} sum of the transverse momenta of partons generated at the hardest interaction of the event, excluding initial- and final-state radiation, i.e. $2 \times$\ttt{PARI(17)}. \iteme{PARI(66) :} sum of the transverse momenta of all partons generated at the hardest interaction, including initial- and final-state radiation, resonance decay products, and primordial $k_{\perp}$. \iteme{PARI(67) :} sum of transverse momenta of partons generated at hard interactions, excluding the hardest one (see \ttt{PARI(65)}), and also excluding initial- and final-state radiation. Is non-vanishing only in the multiple-interaction scenario. \iteme{PARI(68) :} sum of transverse momenta of all partons generated at hard interactions, excluding the hardest one (see \ttt{PARI(66)}), but including initial- and final-state radiation. Is non-vanishing only in the multiple-interaction scenario. \iteme{PARI(69) :} sum of transverse momenta of all partons generated in hard interactions (\ttt{PARI(66) + PARI(68)}) and, additionally, of all beam remnant partons. \iteme{PARI(71), PARI(72) :} sum of the momentum fractions $x$ taken by initial-state parton-shower initiators on side 1 and and side 2, excluding those of the hardest interaction. Is non-vanishing only in the multiple-interaction scenario. \iteme{PARI(73), PARI(74) :} sum of the momentum fractions $x$ taken by the partons at the hard interaction on side 1 and side 2, excluding those of the hardest interaction. Is non-vanishing only in the multiple-interaction scenario. \iteme{PARI(75), PARI(76) :} the $x$ value of a photon that branches into quarks or gluons, i.e. $x$ at interface between initial-state QED and QCD cascades. \iteme{PARI(77), PARI(78) :} the $\chi$ values selected for beam remnants that are split into two objects, describing how the energy is shared (see \ttt{MSTP(92)} and \ttt{MSTP(94)}); is vanishing if no splitting is needed. \iteme{PARI(81) :} size of the threshold factor (enhancement or suppression) in the latest event with heavy-flavour production; see \ttt{MSTP(35)}. \iteme{PARI(91) :} average multiplicity $\br{n}$ of pile-up events, see \ttt{MSTP(133)}. Only relevant for \ttt{MSTP(133)=} 1 or 2. \iteme{PARI(92) :} average multiplicity $\langle n \rangle$ of pile-up events as actually simulated, i.e. with multiplicity = 0 events removed and the high-end tail truncated. Only relevant for \ttt{MSTP(133)=} 1 or 2. \iteme{PARI(93) :} for \ttt{MSTP(133)=1} it is the probability that a beam crossing will produce a pile-up event at all, i.e. that there will be at least one hadron--hadron interaction; for \ttt{MSTP(133)=2} the probability that a beam crossing will produce a pile-up event with one hadron--hadron interaction of the desired rare type. \end{entry} \subsection{How to include external processes in PYTHIA} \label{ss:PYnewproc} Despite a large repertory of processes in {\tsc{Pythia}}, the number of missing ones clearly is even larger, and with time this discrepancy is likely to increase. There are several reasons why it is not practicable to imagine a {\tsc{Pythia}} which has `everything'. One is the amount of time it takes to implement a process for the single {\tsc{Pythia}} author, compared with the rate of new cross section results produced by the rather larger matrix-element calculations community. Another is the length of currently produced matrix-element expressions, which would make the program very bulky. A third argument is that, whereas the phase space of $2 \to 1$ and $2 \to 2$ processes can be set up once and for all according to a reasonably flexible machinery, processes with more final-state particles are less easy to generate. To achieve a reasonable efficiency, it is necessary to tailor the phase-space selection procedure to the dynamics of the given process, and to the desired experimental cuts. If the desired subprocess is missing, it can be included into {\tsc{Pythia}} as an `external' subprocess. In this section we will describe how it is possible to specify the partonic state of some hard-scattering process in an interface common block. {\tsc{Pythia}} will read this common block, and add initial- and final-state showers, beam remnants and underlying events, fragmentation and decays, to build up an event in as much detail as an ordinary {\tsc{Pythia}} one. You may also use {\tsc{Pythia}} to mix events of different kinds, and to keep track of cross section statistics. You have to provide the matrix elements, the phase-space generator, and the storage of event information in the common block. First a minor comment, however. Some processes may be seen just as trivial modifications of already existing ones. For instance, you might want to add some extra term, corresponding to contact interactions, to the matrix elements of a {\tsc{Pythia}} $2 \to 2$ process. In that case it is not necessary to go through the machinery below, but instead you can use the \ttt{PYEVWT} routine to introduce an additional weight for the event, defined as the ratio of the modified to the unmodified differential cross sections. If you use the option \ttt{MSTP(142)=2}, this weight is considered as part of the `true' cross section of the process, and the generation is changed accordingly. The more generic facility for including an external process is a bit more complicated, and involves two routines and one common block. All names contain \ttt{UP}, which is short for User Process. If you want to include a new process, first you have to pick an unused subprocess number ISUB (see tables in section \ref{ss:ISUBcode}). For instance, the numbers 191--200 are currently unused, so this might be a logical place to put a new process. This number and the `title' of the process (plus \ttt{SIGMAX}, to be described below) have to be given in to {\tsc{Pythia}} in a subroutine call \drawbox{CALL PYUPIN(ISUB,TITLE,SIGMAX)}\label{p:PYUPIN} \boxsep \noindent before the call to \ttt{PYINIT}. The \ttt{TITLE} can be any character string up to 28 characters, e.g. \begin{verbatim} CALL PYUPIN(191,'g + g -> t + tbar + gamma',SIGMAX) \end{verbatim} The call to \ttt{PYUPIN} tells the program that a process ISUB exists, but not that you want to generate it. This is done, as with normal processes, by setting \ttt{MSUB(ISUB)=1} before the \ttt{PYINIT} call. Once the event generation chain has been started and \ttt{PYEVNT} is called to generate an event, this routine may in its turn call the routine \ttt{PYUPEV}, which is the routine you must supply, in which the next event is selected. (A dummy copy of \ttt{PYUPEV} has been included at the end of {\tsc{Pythia}}; depending on the machine you may have to comment out this copy when you link your own.) The call arguments are \drawbox{CALL PYUPEV(ISUB,SIGEV)}\label{p:PYUPEV} \boxsep \noindent where \ttt{ISUB} is given by \ttt{PYEVNT}, while \ttt{SIGEV} is to be calculated (see below) and returned to \ttt{PYEVNT}. If there is only one user-defined process, then the \ttt{ISUB} input is superfluous; otherwise it is necessary to branch to the relevant process. The \ttt{SIGEV} variable is supposed to give the differential cross section ot the current event, times the phase-space volume within which events are generated, expressed in millibarns. This means that, in the limit that many events are generated, the average value of \ttt{SIGEV} gives the total cross section of the simulated process. The \ttt{SIGMAX} value, handed to {\tsc{Pythia}} in the \ttt{PYUPIN} call, is assumed to be the maximum value that \ttt{SIGEV} will reach. Events will be accepted with a probability \ttt{SIGEV/SIGMAX}, i.e. the acceptance/rejection of events according to differential cross section is done by \ttt{PYEVNT}, not by the user. This means that the events that come out in the end all have unit weight, i.e. the user does not have to worry about events with different weights. It also allows several subprocesses to be generated together, in the proper mixture. Of course, the tricky part is that the differential cross section usually is strongly peaked in a few regions of the phase space, such that the average probability to accept an event, $\langle$\ttt{SIGEV}$\rangle /$\ttt{SIGMAX} is small. It may then be necessary to find a suitable set of transformed phase-space coordinates, for which the correspondingly transformed differential cross section is better behaved. To avoid unclarities, here is a more formal version of the two above paragraphs. Call $\d X$ the differential phase space, e.g. for a $2 \to 2$ process $\d X = \d x_1 \, \d x_2 \, \d \hat{t}$, where $x_1$ and $x_2$ are the momentum fractions carried by the two incoming partons and $\hat{t}$ the Mandelstam variable of the scattering. Call $\d \sigma / \d X$ the differential cross section of the process, e.g. for $2 \to 2$: $\d \sigma / \d X = \sum_{ij} f_i(x_1,Q^2) \, f_j(x_2,Q^2) \, \d \hat{\sigma}_{ij} / \d \hat{t}$, i.e. the product of parton distributions and hard-scattering matrix elements, summed over all allowed incoming flavours $i$ and $j$. The physical cross section that one then wants to generate is $\sigma = \int (\d \sigma / \d X) \, \d X$, where the integral is over the allowed phase-space volume. The event generation procedure consists of selecting an $X$ uniformly in $\d X$ and then evaluating the weight $\d \sigma / \d X$ at this point. \ttt{SIGEV} is now simply \ttt{SIGEV}$ = \d \sigma / \d X \, \int \d X$, i.e. the differential cross section times the considered volume of phase space. Clearly, when averaged over many events, \ttt{SIGEV} will correctly estimate the desired cross section. If \ttt{SIGEV} fluctuates too much, one may try to transform to new variables $X'$, where events are now picked accordingly to $\d X'$ and \ttt{SIGEV}$ = \d \sigma / \d X' \, \int \d X'$. A warning. It is important that $X$ is indeed uniformly picked within the allowed phase space, alternatively that any Jacobians are properly taken into account. For instance, in the case above, one approach would be to pick $x_1$, $x_2$ and $\hat{t}$ uniformly in the ranges $0 < x_1 < 1$, $0 < x_2 < 1$, and $-s < \hat{t} < 0$, with full phase space volume $\int \d X = s$. The cross section would only be non-vanishing inside the physical region given by $-s x_1 x_2 < \hat{t}$ (in the massless case), i.e. Monte Carlo efficiency is likely to be low. However, if one were to choose $\hat{t}$ values only in the range $-\hat{s} < \hat{t} < 0$, small $\hat{s}$ values would be favoured, since the density of selected $\hat{t}$ values would be larger there. Without the use of a compensating Jacobian $\hat{s}/s$, an incorrect answer would be obtained. Alternatively, one could start out with a phase space like $\d X = \d x_1 \, \d x_2 \, \d (\cos\hat{\theta})$, where the limits decouple. Of course, the $\cos\hat{\theta}$ variable can be translated back into a $\hat{t}$, which will then always be in the desired range $-\hat{s} < \hat{t} < 0$. The transformation itself here gives the necessary Jacobian. If you do not know how big \ttt{SIGMAX} is, you can put it to some very small value (but larger than zero, however) and do an exploratory run. When the program encounters events with \ttt{SIGEV}$ > $\ttt{SIGMAX}, a warning message is printed, which gives the new \ttt{SIGMAX} that the program will use from then on. Hopefully such maximum violations only appear at the beginning of the run, and later \ttt{SIGMAX} stabilizes to a level that can then be used as input for a second, correct run. If you want to do the event rejection yourself, simply put \ttt{SIGEV} equal to \ttt{SIGMAX}. In that case events will not be rejected by {\tsc{Pythia}} (except if there is something else wrong with them). If \ttt{SIGMAX} is the correct total cross section of the process, event mixing with other processes will still work fine. You could also decide not to reject any events, but to use weighted ones. In that case you can only have one ISUB switched on in a run, since the program will not know how to mix different kinds of events, and you cannot use {\tsc{Pythia}} to do cross section statistics for you. Therefore you could, for instance, put \ttt{SIGMAX = SIGEV = 1}, and use a common block to transfer event weight and other information from your \ttt{PYUPEV} routine to your main program. In addition to the \ttt{SIGEV} value returned for each event, it is also necessary to return the event itself. This is done via the common block \drawbox{COMMON/PYUPPR/NUP,KUP(20,7),PUP(20,5),NFUP,IFUP(10,2),% Q2UP(0:10)}\label{p:PYUPPR} \boxsep The first part closely parallels the standard event record in the \ttt{LUJETS} common block, see section \ref{ss:evrec}, although with a few simplifications. The number \ttt{NUP}\label{p:NUP} gives the number of particles involved in the process, where a particle may be a quark, a lepton, a gauge boson, or anything else. The first two are simply the two incoming particles that initiate the hard scattering, while the remaining \ttt{NUP-2} are the outgoing particles from the hard process. For each particle \ttt{I}, with $1 \leq $\ttt{I}$ \leq $\ttt{NUP}, the following information is stored: \begin{entry} \iteme{KUP(I,1) :}\label{p:KUP} is = 1 normally. However, if you put it = 2 that signifies intermediate states that are not to be treated by {\tsc{Pythia}}, but are included only to make the event record easier to read. \iteme{KUP(I,2) :} is the flavour code of a particle, i.e. the two incoming partons for \ttt{I} = 1 and 2, and the outgoing particles for \ttt{I}$ \geq 3$. The flavour codes are the standard KF ones, as used elsewhere in the program. \iteme{KUP(I,3) :} may be used to indicate the position of a mother. Such information may again make the record more readable, but is not really needed, and so one may well put all \ttt{KUP(I,3)=0}. \iteme{KUP(I,4) :} for a final-state parton which carries colour, \ttt{KUP(I,4)} gives the position of the parton from which the colour comes; otherwise it must be 0. \iteme{KUP(I,5) :} for a final-state parton that carries anticolour, \ttt{KUP(I,5)} gives the position of the parton from which the anticolour comes; else it must be 0. \iteme{KUP(I,6) :} for an initial-state parton that carries colour, \ttt{KUP(I,6)} gives the position of the parton to which the colour goes; else it must be 0. \iteme{KUP(I,7) :} for an initial-state parton which carries anticolour, \ttt{KUP(I,7)} gives the position of the parton to which the anticolour goes; else it must be 0. \iteme{PUP(I,1) :}\label{p:PUP} $p_x$, i.e. $x$ momentum. \iteme{PUP(I,2) :} $p_y$, i.e. $y$ momentum. \iteme{PUP(I,3) :} $p_z$, i.e. $z$ momentum. \iteme{PUP(I,4) :} $E$, i.e. energy. \iteme{PUP(I,5) :} $m$, i.e. mass. \end{entry} After this brief summary, we proceed with more details and examples. To illustrate the issue of documentation in \ttt{KUP(I,1)} and \ttt{KUP(I,3)}, consider the case of $\mathrm{W}^+$ production and decay to $\u \overline{\mathrm{d}}$, maybe as part of a more complex process. The final-state particles clearly are $\u$ and $\overline{\mathrm{d}}$, so the $\mathrm{W}^+$ need not be given at all, but it would make the event listing easier to read. Therefore one should add the $\mathrm{W}^+$, but with \ttt{KUP(I,1)=2}. (If the $\mathrm{W}^+$ would have been added with \ttt{KUP(I,1)=1}, it would later have been treated by {\tsc{Pythia}}/{\tsc{Jetset}}, which means it would have been allowed to decay once more.) If the $\mathrm{W}^+$ is in line 3, the $\u$ in 4 and the $\overline{\mathrm{d}}$ in 5, one could further put \ttt{KUP(4,3)=3} and \ttt{KUP(5,3)=3} to indicate that the $\u$ and $\overline{\mathrm{d}}$ in lines 4 and 5 come from the $\mathrm{W}^+$ stored in line 3. The switch \ttt{MSTP(128)} works in the same way for user-defined processes as for ordinary ones, i.e. decay products of resonances can optionally be omitted from the documentation section of the event record, and history pointers can be set slightly differently. The information the program has at its disposal for this purpose is in \ttt{KUP(I,3)}; an entry with this value non-zero is considered as a resonance decay product. The colour-flow information for coloured particles (quarks, gluons, leptoquarks, \ldots) is needed to set up parton showers and fragmentation properly. Sometimes many different colour flows are possible for one and the same process, as discussed in section \ref{sss:QCDjetclass}. It is up to you whether or not you will include all possible colour flows in the appropriate mixture, but at least you must pick some representative colour configuration. Consider e.g. the case of $\mathrm{g}(1) + \mathrm{g}(2) \to \mathrm{q}(3) + \overline{\mathrm{q}}(4)$, where the numbers give the position in the array. It is clear the $\mathrm{q}$ must get its colour from either of the two gluons, which means there are (at least) two possibilities. Picking the $\mathrm{q}$ colour to come from gluon 1, one would thus write \ttt{KUP(3,4)=1}, to be read `the colour of parton 3 comes from parton 1'. By implication therefore also \ttt{KUP(1,6)=3}, i.e. `the colour of parton 1 goes to parton 3', i.e. the colour flow is bookkept doubly. The anticolour now must flow from parton 2 to parton 4, i.e. \ttt{KUP(2,7)=4} and \ttt{KUP(4,5)=2}. This completely specifies the colours of the $\mathrm{q}$ and $\overline{\mathrm{q}}$, but not of the two gluons. In fact, one colour in the initial state `annihilates' between the $\mathrm{g}(1)$ and $\mathrm{g}(2)$, i.e. the anticolour of gluon 1 and the colour of gluon 2 match, which may be expressed by \ttt{KUP(1,7)=2} and \ttt{KUP(2,6)=1}. In other words colour/anticolour of an initial-state parton may either go to a final-state parton or to another initial-state parton. Correspondingly, the colour/anticolour of a final-state parton may come either from an initial-state parton or from another final-state parton. An example of the latter possibility is $\mathrm{W}$ decays, or generically the decay of any colour-singlet particle. (Thus a third colour flow above is represented by $\mathrm{g} \mathrm{g} \to \H^0 \to \mathrm{q} \overline{\mathrm{q}}$, where no colour passes through the Higgs, and therefore colour flows between the two gluons and, separately, between the $\mathrm{q}$ and $\overline{\mathrm{q}}$.) Storing of momenta should be straightforward, but a few comments must be made. Even if you ask, in the \ttt{PYINIT} call, to have events generated in a fixed target or a user-specified frame, at intermediate stages {\tsc{Pythia}} will still work in the c.m. frame of the two incoming beam particles, with the first beam moving in the $+z$ direction and the second in the $-z$ one. This c.m. frame must also be used when giving the momenta of the process. In addition, the two incoming partons in lines 1 and 2 are assumed massless. Therefore the initial-state partons are characterized only by the two energies \ttt{P(1,4)} and \ttt{P(2,4)}, with \ttt{P(1,3) = P(1,4)}, \ttt{P(2,3) = -P(2,4)}, and everything else is zero. In the final state, energies, momenta and masses are free, but must add up to give the same four-momentum as that of the initial state. All momenta are given in GeV, with the speed of light $c = 1$. The second part of the \ttt{PYUPPR} common block is used to regulate the initial- and final-state showering, as follows: \begin{entry} \iteme{Q2UP(0) :}\label{p:Q2UP} $Q^2$ scale of initial-state showers. \iteme{NFUP :}\label{p:NFUP} number of parton pairs that undergo final-state showers. \iteme{IFUP(IF,1), IFUP(IF,2) :}\label{p:IFUP} positions of the two partons of a final-state showering pair, where the index \ttt{IF} runs between 1 and \ttt{NFUP}. \iteme{Q2UP(IF) :} the $Q^2$ scale of the final-state shower between parton pair \ttt{IF} above. \end{entry} If you do not want any showering at all, you can put \ttt{MSTP(61)=0} and \ttt{MSTP(71)=0}, and then you do not have to give the above quantities. In general the scale choices $Q^2$ are not unique, which means that some guesswork is involved. Since the showers add extra partonic activity at mass scales below the mentioned $Q^2$ choices, the \ttt{Q2UP} should be of the order of the phase-space cut-offs, so as to provide a reasonably smooth joining between partonic activity from matrix elements and that from showers. There are a few cases where choices are rather easy. In the decay of any $s$-channel colour neutral state, such as a $\mathrm{W}^{\pm}$, the $Q^2$ scale of final-state showers is just set by the squared mass of the resonance. For initial-state radiation, \ttt{Q2UP(0)} should be about the same as the $Q^2$ scale used for the evaluation of parton distributions for the hard process, up to some factor of order unity. (One frequent choice for this factor would be 4, if your parton-distribution scale is something like the squared transverse momentum, simply because $m^2$ is of order $4 p_{\perp}^2$.) The `parton'-shower evolution actually also can include photon emission off quarks and leptons, if the shower switches are properly set. It is not possible to define only one particle in the above arrays, since it would then not be possible to conserve energy and momentum in the shower. You can very well have a pair where only one of the two can branch, however. For instance, in a $\mathrm{g} \gamma$ final state, only the gluon can shower, but the photon can lose energy to the gluon in such a way that the gluon branchings becomes possible. Currently, it is not possible to do showering where three or more final-state particles are involved at the same time. This may be added at a later stage. It is therefore necessary to subdivide suitably into pairs, and maybe leave some (especially colour-neutral) particles unshowered. You are free to make use of whatever tools you want in your \ttt{PYUPEV} routine, and normally there would be no contact with the rest of {\tsc{Pythia}}, except as described above. However, you may want to use some of the tools already available. One attractive possibility is to use \ttt{PYSTFU} for parton-distribution-function evaluation. Other possible tools could be \ttt{RLU} for random-number generation, \ttt{ULALPS} for $\alpha_{\mathrm{s}}$ evaluation, \ttt{ULALEM} for evaluation of a running $\alpha_{\mathrm{em}}$, and maybe a few more. We end with a few comments on anomalous situations. In some cases one may want to decide, inside \ttt{PYUPEV}, when to stop the event-generation loop. This is the case, for instance, if event configurations are read in from a file, and the end of the file is reached. One might be tempted just to put \ttt{SIGEV=0} when this happens. Then \ttt{PYEVNT} will discard the event, as part of the matrix-element-weighting procedure. However, next \ttt{PYEVNT} will generate another event, which normally means a new request to \ttt{PYUPEV}, so one does not really get out of the loop. Instead you should put \ttt{NUP=0}. If the program encounters this value at a return from \ttt{PYUPEV}, then it will also exit from \ttt{PYEVNT}, without incrementing the counters for the number of events generated. It is then up to you to have a check on this condition in your main event-generation loop. This you do either by looking at \ttt{NUP} or at \ttt{MSTI(51)}; the latter is set to 1 if no event was generated. It may also happen that a user-defined configuration fails elsewhere in the \ttt{PYEVNT} call. For instance, the beam-remnant treatment occasionally encounters situations it cannot handle, wherefore the hard interaction is rejected and a new one generated. This happens also with ordinary (not user-defined) events, and usually comes about as a consequence of the initial-state radiation description leaving too little energy for the remnant. If the same hard scattering were to be used as input for a new initial-state radiation and beam-remnant attempt, it could then work fine. There is a possibility to give events that chance, as follows. \ttt{MSTI(52)} counts the number of times a hard-scattering configuration has failed to date. If you come in to \ttt{PYUPEV} with \ttt{MSTI(52)} non-vanishing, this means that the latest configuration failed. So long as the contents of the \ttt{PYUPPR} common block are not changed, such an event may be given another try. For instance, a line \begin{verbatim} IF(MSTI(52).GE.1.AND.MSTI(52).LE.4) RETURN \end{verbatim} at the beginning of \ttt{PYUPEV} will give each event up to five tries; thereafter a new one would be generated as usual. Note that the counter for the number of events is updated at each new try. The fraction of failed configurations is given in the bottom line of the \ttt{PYSTAT(1)} table. The above comment only refers to very rare occurrences (less than one in a hundred), which are not errors in a strict sense; for instance, they do not produce any error messages on output. If you get warnings and error messages that the program does not understand the flavour codes or cannot reconstruct the colour flows, it is due to faults of yours, and giving such events more tries is not going to help. \subsection{How to run PYTHIA with varying energies} \label{ss:PYvaren} It is possible to use {\tsc{Pythia}} in a mode where the energy can be varied from one event to the next, without the need to reinitialize with a new \ttt{PYINIT} call. This allows a significant speed-up of execution, although it is not as fast as running at a fixed energy. It can not be used for everything --- we will come to the fine print at the end --- but it should be applicable for most tasks. The master switch to access this possibility is in \ttt{MSTP(171)}. By default it is off, so you must set \ttt{MSTP(171)=1} before initialization. There are two submodes of running, with \ttt{MSTP(172)} being 1 or 2. In the former mode, {\tsc{Pythia}} will generate an event at the requested energy. This means that you have to know which energy you want beforehand. In the latter mode, {\tsc{Pythia}} will often return without having generated an event --- with flag \ttt{MSTI(61)=1} to signal that --- and you are then requested to give a new energy. The energy spectrum of accepted events will then, in the end, be your naive input spectrum weighted with the cross-section of the processes you study. We will come back to this. The energy can be varied, whichever frame is given in the \ttt{PYINIT} call. When the frame is \ttt{'CMS'}, \ttt{PARP(171)} should be filled with the fractional energy of each event, i.e. $E_{\mrm{cm}} =$\ttt{PARP(171)}$\times$\ttt{WIN}, where \ttt{WIN} is the nominal c.m. energy of the \ttt{PYINIT} call. Here \ttt{PARP(171)} should normally be smaller than unity, i.e. initialization should be done at the maximum energy to be encountered. For the \ttt{'FIXT'} frame, \ttt{PARP(171)} should be filled by the fractional beam energy of that one, i.e. $E_{\mrm{beam}} = $\ttt{PARP(171)}$\times$\ttt{WIN}. For the \ttt{'USER'}, \ttt{'FOUR'} and \ttt{'FIVE'} options, the two four-momenta are given in for each event in the same format as used for the \ttt{PYINIT} call. Note that there is a minimum c.m. energy allowed, \ttt{PARP(2)}. If you give in values below this, the program will stop for \ttt{MSTP(172)=1}, and will return with \ttt{MSTI(61)=1} for \ttt{MSTP(172)=1}. To illustrate the use of the \ttt{MSTP(172)=2} facility, consider the case of beamstrahlung in $\e^+\e^-$ linear colliders. This is just for convenience; what is said here can be translated easily into other situations. Assume that the beam spectrum is given by $D(z)$, where $z$ is the fraction retained by the original $\mathrm{e}$ after beamstrahlung. Therefore $0 \leq z \leq 1$ and the integral of $D(z)$ is unity. This is not perfectly general; one could imagine branchings $\mathrm{e}^- \to \mathrm{e}^- \gamma \to \mathrm{e}^-\mathrm{e}^+\mathrm{e}^-$, which gives a multiplication in the number of beam particles. This could either be expressed in terms of a $D(z)$ with integral larger than unity or in terms of an increased luminosity. We will assume the latter, and use $D(z)$ properly normalized. Given a nominal $s = 4E_{\mrm{beam}}^2$, the actual $s'$ after beamstrahlung is given by $s' = z_1 z_2 s$. For a process with a cross section $\sigma(s)$ the total cross section is then \begin{equation} \sigma_{\mrm{tot}} = \int_0^1 \int_0^1 D(z_1) \, D(z_2) \sigma(z_1 z_2 s) \, \d z_1 \, \d z_2~. \end{equation} The cross section $\sigma$ may in itself be an integral over a number of additional phase space variables. If the maximum of the differential cross section is known, a correct procedure to generate events is \begin{Enumerate} \item pick $z_1$ and $z_2$ according to $D(z_1) \, \d z_1$ and $D(z_2) \, \d z_2$, respectively; \item pick a set of phase space variables of the process, for the given $s'$ of the event; \item evaluate $\sigma(s')$ and compare with $\sigma_{\mathrm{max}}$; \item if event is rejected, then return to step 1 to generate new variables; \item else continue the generation to give a complete event. \end{Enumerate} You as a user are assumed to take care of step 1, and present the resulting kinematics with incoming $\mathrm{e}^+$ and $\mathrm{e}^-$ of varying energy. Thereafter {\tsc{Pythia}} will do steps 2--5, and either return an event or put \ttt{MSTI(61)=1} to signal failure in step 4. The maximization procedure does search in phase space to find $\sigma_{\mathrm{max}}$, but it does not vary the $s'$ energy in this process. Therefore the maximum search in the \ttt{PYINIT} call should be performed where the cross section is largest. For processes with increasing cross section as a function of energy this means at the largest energy that will ever be encountered, i.e. $s' = s$ in the case above. This is the `standard' case, but often one encounters other behaviours, where more complicated procedures are needed. One such case would be the process $\e^+\e^- \to \mathrm{Z}^{*0} \to \mathrm{Z}^0 \H^0$, which is known to have a cross section that increases near the threshold but is decreasing asymptotically. If one already knows that the maximum, for a given Higgs mass, appears at 300 GeV, say, then the \ttt{PYINIT} call should be made with that energy, even if subsequently one will be generating events for a 500 GeV collider. In general, it may be necessary to modify the selection of $z_1$ and $z_2$ and assign a compensating event weight. For instance, consider a process with a cross section behaving roughly like $1/s$. Then the $\sigma_{\mrm{tot}}$ expression above may be rewritten as \begin{equation} \sigma_{\mrm{tot}} = \int_0^1 \int_0^1 \frac{D(z_1)}{z_1} \, \frac{D(z_2)}{z_2} \, z_1 z_2 \sigma(z_1 z_2 s) \, \d z_1 \, \d z_2 ~. \end{equation} The expression $z_1 z_2 \sigma(s')$ is now essentially flat in $s'$, i.e. not only can $\sigma_{\mathrm{max}}$ be found at a convenient energy such as the maximum one, but additionally the {\tsc{Pythia}} generation efficiency (the likelihood of surviving step 4) is greatly enhanced. The price to be paid is that $z$ has to be selected according to $D(z)/z$ rather than according to $D(z)$. Note that $D(z)/z$ is not normalized to unity. One therefore needs to define \begin{equation} {\cal I}_D = \int_0^1 \frac{D(z)}{z} \, \d z ~, \end{equation} and a properly normalized \begin{equation} D'(z) = \frac{1}{{\cal I}_D} \, \frac{D(z)}{z} ~. \end{equation} Then \begin{equation} \sigma_{\mrm{tot}} = \int_0^1 \int_0^1 D'(z_1) \, D'(z_2) \, {\cal I}_D^2 \, z_1 z_2 \sigma(z_1 z_2 s) \, \d z_1 \, \d z_2 ~. \end{equation} Therefore the proper event weight is ${\cal I}_D^2 \, z_1 z_2$. This weight should be stored, for each event, in \ttt{PARP(173)}. The maximum weight that will be encountered should be stored in \ttt{PARP(174)} before the \ttt{PYINIT} call, and not changed afterwards. It is not necessary to know the precise maximum; any value larger than the true maximum will do, but the inefficiency will be larger the cruder the approximation. Additionally you must put \ttt{MSTP(173)=1} for the program to make use of weights at all. Often $D(z)$ are not known analytically; therefore ${\cal I}_D$ is also not known beforehand, but may have to be evaluated (by you) during the course of the run. Then you should just use the weight $z_1 z_2$ in \ttt{PARP(173)} and do the overall normalization yourself in the end. Since \ttt{PARP(174)=1.} by default, in this case you need not set this variable specially. Only the cross sections are affected by the procedure selected for overall normalization, the events themselves still are properly distributed in $s'$ and internal phase space. Above it has been assumed tacitly that $D(z) \to 0$ for $z \to 0$. If not, $D(z)/z$ is divergent, and it is not possible to define a properly normalized $D'(z) = D(z)/z$. If the cross section is truly diverging like $1/s$, then a $D(z)$ which is nonvanishing for $z \to 0$ does imply an infinite total cross section, whichever way things are considered. In cases like that, it is necessary to impose a lower cut on $z$, based on some physics or detector consideration. Some such cut is anyway needed to keep away from the minimum c.m. energy required for {\tsc{Pythia}} events, see above. The most difficult cases are those with a very narrow and high peak, such as the $\mathrm{Z}^0$. One could initialize at the energy of maximum cross section and use $D(z)$ as is, but efficiency might turn out to be very low. One might then be tempted to do more complicated transforms of the kind illustrated above. As a rule it is then convenient to work in the variables $\tau_z = z_1 z_2$ and $y_z = (1/2) \ln (z_1/z_2)$, cf. subsection \ref{ss:kinemtwo}. Clearly, the better the behaviour of the cross section can be modelled in the choice of $z_1$ and $z_2$, the better the overall event generation efficieny. Even under the best of circumstances, the efficiency will still be lower than for runs with fix energy. There is also a non-negligible time overhead for using variable energies in the first place, from kinematics reconstruction and (in part) from the phase space selection. One should therefore not use variable energies when not needed, and not use a large range of energies $\sqrt{s'}$ if in the end only a smaller range is of experimental interest. This facility may be combined with most other aspects of the program. For instance, it is possible to simulate beamstrahlung as above and still include bremsstrahlung with \ttt{MSTP(11)=1}. Further, one may multiply the overall event weight of \ttt{PARP(173)} with a kinematics-dependent weight given by \ttt{PYEVWT}, although it is not recommended (since the chances of making a mistake are also multiplied). However, a few things do \textit{not} work. \begin{Itemize} \item It is not possible to use pile-up events, i.e. you must have \ttt{MSTP(131)=0}. \item The possibility of giving in your own cross-section optimization coefficients, option \ttt{MSTP(121)=2}, would require more input than with fixed energies, and this option should therefore not be used. You can still use \ttt{MSTP(121)=1}, however. \item The multiple interactions scenario with \ttt{MSTP(82)}$\geq 2$ only works approximately for energies different from the initialization one. If the c.m. energy spread is smaller than a factor 2, say, the approximation should be reasonable, but if the spread is larger one may have to subdivide into subruns of different energy bins. The initialization should be made at the largest energy to be encountered --- whenever multiple interactions are possible (i.e. for incoming hadrons and resolved photons) this is where the cross sections are largest anyway, and so this is no further constraint. There is no simple possibility to change \ttt{PARP(82)} during the course of the run, i.e. an energy-independent $p_{\perp 0}$ must be assumed. The default option \ttt{MSTP(82)=1} works fine, i.e. does not suffer from the constraints above. If so desired, $p_{\perp\mathrm{min}}=$\ttt{PARP(81)} can be set differently for each event, as a function of c.m. energy. Initialization should then be done with \ttt{PARP(81)} as low as it is ever supposed to become. \end{Itemize} \subsection{Other Routines and Common Blocks} The subroutines and common blocks that you will come in direct contact with have already been described. A number of other routines and common blocks exist, and those not described elsewhere are here briefly listed for the sake of completeness. The \ttt{PYG***} routines are slightly modified versions of the \ttt{SAS***} ones of the \tsc{SaSgam} library. The common block \ttt{SASCOM} is renamed \ttt{PYINT8}. If you want to use the parton distributions for standalone purposes, you are encouraged to use the original \tsc{SaSgam} routines rather than going the way via the {\tsc{Pythia}} adaptations. \begin{entry} \iteme{SUBROUTINE PYINRE :}\label{p:PYINRE} to initialize the widths and effective widths of resonances. \iteme{SUBROUTINE PYINBM(CHFRAM,CHBEAM,CHTARG,WIN) :}\label{p:PYINBM} to read in and identify the beam (\ttt{CHBEAM}) and target (\ttt{CTTARG}) particles and the frame (\ttt{CHFRAM}) as given in the \ttt{PYINIT} call; also to save the original energy (\ttt{WIN}). \iteme{SUBROUTINE PYINKI(MODKI) :}\label{p:PYINKI} to set up the event kinematics, either at initialization (\ttt{MODKI=0}) or for each separate event, the latter when the program is run with varying kinematics (\ttt{MODKI=1}). \iteme{SUBROUTINE PYINPR :}\label{p:PYINPR} to set up the partonic subprocesses selected with \ttt{MSEL}. For $\gamma\mathrm{p}$ and $\gamma\gamma$, also the \ttt{MSTP(14)} value affects the choice of processes. In particular, the option \ttt{MSTP(14)=10} sets up the three or six different processes that need to be mixed, with separate cuts for each. \iteme{SUBROUTINE PYXTOT :}\label{p:PYXTOT} to give the parametrized total, double diffractive, single diffractive and elastic cross sections for different energies and colliding hadrons. \iteme{SUBROUTINE PYMAXI :}\label{p:PYMAXI} to find optimal coefficients \ttt{COEF} for the selection of kinematical variables, and to find the related maxima for the differential cross section times Jacobian factors, for each of the subprocesses included. \iteme{SUBROUTINE PYPILE(MPILE) :}\label{p:PYPILE} to determine the number of pile-up events, i.e. events appearing in the same beam--beam crossing. \iteme{SUBROUTINE PYSAVE(ISAVE,IGA) :}\label{p:PYSAVE} saves and restores parameters and cross section values between the three $\gamma\mathrm{p}$ and the six $\gamma\gamma$ components of \ttt{MSTP(14)=10}. The options for \ttt{ISAVE} are (1) a complete save of all parameters specific to a given component, (2) a partial save of cross-section information, (3) a restoration of all parameters specific to a given component, (4) as 3 but preceded by a random selection of component, and (5) a summation of component cross sections (for \ttt{PYSTAT}). The subprocess code in \ttt{IGA} is the one described for \ttt{MSTI(9)}; it is input for options 1, 2 and 3 above, output for 4 and dummy for 5. \iteme{SUBROUTINE PYRAND :}\label{p:PYRAND} to generate the quantities characterizing a hard scattering on the parton level, according to the relevant matrix elements. \iteme{SUBROUTINE PYSCAT :}\label{p:PYSCAT} to find outgoing flavours and to set up the kinematics and colour flow of the hard scattering. \iteme{SUBROUTINE PYRESD :}\label{p:PYRESD} to allow resonances to decay, including chains of successive decays and parton showers. \iteme{SUBROUTINE PYMULT(MMUL) :}\label{p:PYMULT} to generate semi-hard interactions according to the multiple interaction formalism. \iteme{SUBROUTINE PYREMN(IPU1,IPU2) :}\label{p:PYREMN} to add on target remnants and include primordial $k_{\perp}$. \iteme{SUBROUTINE PYDIFF :}\label{p:PYDIFF} to handle diffractive and elastic scattering events. \iteme{SUBROUTINE PYDOCU :}\label{p:PYDOCU} to compute cross sections of processes, based on current Monte Carlo statistics, and to store event information in the \ttt{MSTI} and \ttt{PARI} arrays. \iteme{SUBROUTINE PYWIDT(KFLR,SH,WDTP,WDTE) :}\label{p:PYWIDT} to calculate widths and effective widths of resonances. \iteme{SUBROUTINE PYOFSH(MOFSH,KFMO,KFD1,KFD2,PMMO,RET1,RET2) :}% \label{p:PYOFSH} to calculate partial widths into channels off the mass shell, and to select correlated masses of resonance pairs. \iteme{SUBROUTINE PYKLIM(ILIM) :}\label{p:PYKLIM} to calculate allowed kinematical limits. \iteme{SUBROUTINE PYKMAP(IVAR,MVAR,VVAR) :}\label{p:PYKMAP} to calculate the value of a kinematical variable when this is selected according to one of the simple pieces. \iteme{SUBROUTINE PYSIGH(NCHN,SIGS) :}\label{p:PYSIGH} to give the differential cross section (multiplied by the relevant Jacobians) for a given subprocess and kinematical setup. \iteme{SUBROUTINE PYSTFL(KF,X,Q2,XPQ) :}\label{p:PYSTFL} to give parton distributions for $\mathrm{p}$ and $\mathrm{n}$ in the option with modified behaviour at small $Q^2$ and $x$, see \ttt{MSTP(57)}. \iteme{SUBROUTINE PYSTFU(KF,X,Q2,XPQ) :}\label{p:PYSTFU} to give parton-distribution functions (multiplied by $x$, i.e. $x f_i(x,Q^2)$) for an arbitrary particle (of those recognized by {\tsc{Pythia}}). Generic driver routine for the following, specialized ones. \begin{subentry} \iteme{KF :} flavour of probed particle, according to KF code. \iteme{X :} $x$ value at which to evaluate parton distributions. \iteme{Q2 :} $Q^2$ scale at which to evaluate parton distributions. \iteme{XPQ :} array of dimensions \ttt{XPQ(-25:25)}, which contains the evaluated parton distributions $x f_i(x,Q^2)$. Components $i$ ordered according to standard KF code; additionally the gluon is found in position 0 as well as 21 (for historical reasons). \end{subentry} \iteme{SUBROUTINE PYSTEL(X,Q2,XPEL) :}\label{p:PYSTEL} to give electron parton distributions. \iteme{SUBROUTINE PYSTGA(X,Q2,XPGA) :}\label{p:PYSTGA} to give the photon parton distributions for sets other than the SaS ones. \iteme{SUBROUTINE PYGGAM(ISET,X,Q2,P2,F2GM,XPDFGM) :}\label{p:PYGGAM} to construct the SaS $F_2$ and parton distributions of the photon by summing homogeneous (VMD) and inhomogeneous (anomalous) terms. For $F_2$, $\c$ and $\b$ are included by the Bethe-Heitler formula; in the `{$\overline{\mrm{MS}}$}' scheme additionally a $C^{\gamma}$ term is added. Calls \ttt{PYGVMD}, \ttt{PYGANO}, \ttt{PYGBEH}, and \ttt{PYGDIR}. \iteme{SUBROUTINE PYGVMD(ISET,KF,X,Q2,P2,ALAM,XPGA) :}\label{p:PYGVMD} to evaluate the VMD parton distributions of a photon, evolved homogeneously from an initial scale $P^2$ to $Q^2$. \iteme{SUBROUTINE PYGANO(KF,X,Q2,P2,ALAM,XPGA) :}\label{p:PYGANO} to evaluate the parton distributions of the anomalous photon, inhomogeneously evolved from a scale $P^2$ (where it vanishes) to $Q^2$. \iteme{SUBROUTINE PYGBEH(KF,X,Q2,P2,PM2,XPBH) :}\label{p:PYGBEH} to evaluate the Bethe-Heitler cross section for heavy flavour production. \iteme{SUBROUTINE PYGDIR(X,Q2,P2,AK0,XPGA) :}\label{p:PYGDIR} to evaluate the direct contribution, i.e. the $C^{\gamma}$ term, as needed in {$\overline{\mrm{MS}}$} parametrizations. \iteme{SUBROUTINE PYSTPI(X,Q2,XPPI) :}\label{p:PYSTPI} to give pion parton distributions. \iteme{SUBROUTINE PYSTPR(X,Q2,XPPR) :}\label{p:PYSTPR} to give proton parton distributions. \iteme{FUNCTION PYCTQ2(ISET,IPRT,X,Q) :}\label{p:PYCTQ2} to give the CTEQ2 proton parton distributions. \iteme{FUNCTION PYHFTH(SH,SQM,FRATT) :}\label{p:PYHFTH} to give heavy-flavour threshold factor in matrix elements. \iteme{SUBROUTINE PYSPLI(KF,KFLIN,KFLCH,KFLSP) :}\label{p:PYSPLI} to give hadron remnant or remnants left behind when the reacting parton is kicked out. \iteme{FUNCTION PYGAMM(X) :}\label{p:PYGAMM} to give the value of the ordinary $\Gamma (x)$ function (used in some parton-distribution parametrizations). \iteme{SUBROUTINE PYWAUX(IAUX,EPS,WRE,WIM) :}\label{p:PYWAUX} to evaluate the two auxiliary functions $W_1$ and $W_2$ appearing in the cross section expressions in \ttt{PYSIGH}. \iteme{SUBROUTINE PYI3AU(EPS,RAT,Y3RE,Y3IM) :}\label{p:PYI3AU} to evaluate the auxiliary function $I_3$ appearing in the cross section expressions in \ttt{PYSIGH}. \iteme{FUNCTION PYSPEN(XREIN,XIMIN,IREIM) :}\label{p:PYSPEN} to calculate the real and imaginary part of the Spence function. \iteme{SUBROUTINE PYQQBH(WTQQBH) :}\label{p:PYQQBH} to calculate matrix elements for the two processes $\mathrm{g} \mathrm{g} \to \mathrm{Q} \overline{\mathrm{Q}} \H^0$ and $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{Q} \overline{\mathrm{Q}} \H^0$. \iteme{BLOCK DATA PYDATA :}\label{p:PYDATA} to give sensible default values to all status codes and parameters. \end{entry} \drawbox{COMMON/PYINT1/MINT(400),VINT(400)}\label{p:PYINT1} \begin{entry} \itemc{Purpose:} to collect a host of integer- and real-valued variables used internally in the program during the initialization and/or event generation stage. These variables must not be changed by you. \iteme{MINT(1) :}\label{p:MINT} specifies the general type of subprocess that has occurred, according to the ISUB code given in section \ref{ss:ISUBcode}. \iteme{MINT(2) :} whenever \ttt{MINT(1)} (together with \ttt{MINT(15)} and \ttt{MINT(16)}) are not sufficient to specify the type of process uniquely, \ttt{MINT(2)} provides an ordering of the different possibilities, see \ttt{MSTI(2)}. \iteme{MINT(3) :} number of partons produced in the hard interactions, i.e. the number $n$ of the $2 \to n$ matrix elements used; is sometimes 3 or 4 when a basic $2 \to 1$ or $2 \to 2$ process has been folded with two $1 \to 2$ initial branchings (like $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \mathrm{q}''' \H^0$). \iteme{MINT(4) :} number of documentation lines at the beginning of the common block \ttt{LUJETS} that are given with \ttt{K(I,1)=21}; 0 for \ttt{MSTP(125)=0}. \iteme{MINT(5) :} number of events generated to date in current run. In runs with the variable-energy option, \ttt{MSTP(171)=1} and \ttt{MSTP(172)=2}, only those events that survive (i.e. that do not have \ttt{MSTI(61)=1}) are counted in this number. That is, \ttt{MINT(5)} may be less than the total number of \ttt{PYEVNT} calls. \iteme{MINT(6) :} current frame of event (see \ttt{MSTP(124)} for possible values). \iteme{MINT(7), MINT(8) :} line number for documentation of outgoing partons/particles from hard scattering for $2 \to 2$ or $2 \to 1 \to 2$ processes (else = 0). \iteme{MINT(10) :} is 1 if cross section maximum was violated in current event, and 0 if not. \iteme{MINT(11) :} KF flavour code for beam (side 1) particle. \iteme{MINT(12) :} KF flavour code for target (side 2) particle. \iteme{MINT(13), MINT(14) :} KF flavour codes for side 1 and side 2 initial-state shower initiators. \iteme{MINT(15), MINT(16) :} KF flavour codes for side 1 and side 2 incoming partons to the hard interaction. \iteme{MINT(17), MINT(18) :} flag to signal if particle on side 1 or side 2 has been scattered diffractively; 0 if no, 1 if yes. \iteme{MINT(19), MINT(20) :} flag to signal initial-state structure with parton inside photon inside electron on side 1 or side 2; 0 if no, 1 if yes. \iteme{MINT(21) - MINT(24) :} KF flavour codes for outgoing partons from the hard interaction. The number of positions actually used is process-dependent, see \ttt{MINT(3)}; trailing positions not used are set = 0. \iteme{MINT(25), MINT(26) :} KF flavour codes of the products in the decay of a single $s$-channel resonance formed in the hard interaction. Are thus only used when \ttt{MINT(3)=1} and the resonance is allowed to decay. \iteme{MINT(31) :} number of hard or semi-hard scatterings that occurred in the current event in the multiple-interaction scenario; is = 0 for a low-$p_{\perp}$ event. \iteme{MINT(35) :} in a true $2 \to 3$ process, where one particle is a resonance with decay channel selected already before the \ttt{PYRESD} call, the decay channel number (in the \ttt{/LUDAT3/} numbering) is stored here. \iteme{MINT(41), MINT(42) :} type of incoming beam or target particle; 1 for lepton and 2 for hadron. A photon counts as a lepton if it is not resolved (\ttt{MSTP(14)=0}) and as a hadron if it is resolved (\ttt{MSTP(14)}$\geq 1$). \iteme{MINT(43) :} combination of incoming beam and target particles. A photon counts as a hadron. \begin{subentry} \iteme{= 1 :} lepton on lepton. \iteme{= 2 :} lepton on hadron. \iteme{= 3 :} hadron on lepton. \iteme{= 4 :} hadron on hadron. \end{subentry} \iteme{MINT(44) :} as \ttt{MINT(43)}, but a photon counts as a lepton. \iteme{MINT(45), MINT(46) :} structure of incoming beam and target particles. \begin{subentry} \iteme{= 1 :} no internal structure, i.e. an electron or photon carrying the full beam energy. \iteme{= 2 :} defined with parton distributions that are not peaked at $x = 1$, i.e. a hadron or a resolved photon. \iteme{= 3 :} defined with parton distributions that are peaked at $x = 1$, i.e. a resolved electron. \end{subentry} \iteme{MINT(47) :} combination of incoming beam- and target-particle parton-distribution function types. \begin{subentry} \iteme{= 1 :} no parton distribution either for beam or target. \iteme{= 2 :} parton distributions for target but not for beam. \iteme{= 3 :} parton distributions for beam but not for target. \iteme{= 4 :} parton distributions for both beam and target, but not both peaked at $x = 1$. \iteme{= 5 :} parton distributions for both beam and target, with both peaked at $x = 1$. \end{subentry} \iteme{MINT(48) :} total number of subprocesses switched on. \iteme{MINT(49) :} number of subprocesses that are switched on, apart from elastic scattering and single, double and central diffractive. \iteme{MINT(50) :} combination of incoming particles from a multiple interactions point of view. \begin{subentry} \iteme{= 0 :} the total cross section is not known; therefore no multiple interactions are possible. \iteme{= 1 :} the total cross section is known; therefore multiple interactions are possible if switched on. \end{subentry} \iteme{MINT(51) :} internal flag that event failed cuts. \begin{subentry} \iteme{= 0 :} no problem. \iteme{= 1 :} event failed; new one to be generated. \iteme{= 2 :} event failed; no new event is to be generated but instead control is to be given back to used. Is intended for user-defined processes, when \ttt{NUP=0}. \end{subentry} \iteme{MINT(52) :} internal counter for number of lines used (in \ttt{/LUJETS/}) before multiple interactions are considered. \iteme{MINT(53) :} internal counter for number of lines used (in \ttt{/LUJETS/}) before beam remnants are considered. \iteme{MINT(55) :} the heaviest new flavour switched on for QCD processes, specifically the flavour to be generated for ISUB = 81, 82, 83 or 84. \iteme{MINT(56) :} the heaviest new flavour switched on for QED processes, specifically for ISUB = 85. Note that, unlike \ttt{MINT(55)}, the heaviest flavour may here be a lepton, and that heavy means the one with largest KF code. \iteme{MINT(57) :} number of times the beam remnant treatment has failed, and the same basic kinematical setup is used to produce a new parton shower evolution and beam remnant set. Mainly used in leptoproduction, for the option when $x$ and $Q^2$ are to be preserved. \iteme{MINT(61) :} internal switch for the mode of operation of resonance width calculations in \ttt{PYWIDT} for $\gamma^* / \Z^0$ or $\gamma^*/\mathrm{Z}^0/\mathrm{Z}'^0$. \begin{subentry} \iteme{= 0 :} without reference to initial-state flavours. \iteme{= 1 :} with reference to given initial-state flavours. \iteme{= 2 :} for given final-state flavours. \end{subentry} \iteme{MINT(62) :} internal switch for use at initialization of $\H^0$ width. \begin{subentry} \iteme{= 0 :} use widths into $\mathrm{Z} \mathrm{Z}^*$ or $\mathrm{W} \mathrm{W}^*$ calculated before. \iteme{= 1 :} evaluate widths into $\mathrm{Z} \mathrm{Z}^*$ or $\mathrm{W} \mathrm{W}^*$ for current Higgs mass. \end{subentry} \iteme{MINT(65) :} internal switch to indicate initialization without specified reaction. \begin{subentry} \iteme{= 0 :} normal initialization. \iteme{= 1 :} initialization with argument \ttt{'none'} in \ttt{PYINIT} call. \end{subentry} \iteme{MINT(71) :} switch to tell whether current process is singular for $p_{\perp} \to 0$ or not. \begin{subentry} \iteme{= 0 :} non-singular process, i.e. proceeding via an $s$-channel resonance or with both products having a mass above \ttt{CKIN(6)}. \iteme{= 1 :} singular process. \end{subentry} \iteme{MINT(72) :} number of $s$-channel resonances that may contribute to the cross section. \iteme{MINT(73) :} KF code of first $s$-channel resonance; 0 if there is none. \iteme{MINT(74) :} KF code of second $s$-channel resonance; 0 if there is none. \iteme{MINT(81) :} number of selected pile-up events. \iteme{MINT(82) :} sequence number of currently considered pile-up event. \iteme{MINT(83) :} number of lines in the event record already filled by previously considered pile-up events. \iteme{MINT(84) :} \ttt{MINT(83) + MSTP(126)}, i.e. number of lines already filled by previously considered events plus number of lines to be kept free for event documentation. \iteme{MINT(91) :} is 1 for a lepton--hadron event and 0 else. Used to determine whether a \ttt{PYFRAM(3)} call is possible. \iteme{MINT(92) :} is used to denote region in $(x,Q^2)$ plane when \ttt{MSTP(57)=2}, according to numbering in \cite{Sch93a}. Simply put, 0 means that the modified proton parton distributions were not used, 1 large $x$ and $Q^2$, 2 small $Q^2$ but large $x$, 3 small $x$ but large $Q^2$ and 4 small $x$ and $Q^2$. \iteme{MINT(93) :} is used to keep track of parton distribution set used in the latest \ttt{STRUCTM} call to \tsc{Pdflib}. The code for this set is stored in the form \ttt{MINT(93) = 1000000}$\times$\ttt{NPTYPE + 1000}$\times$% \ttt{NGROUP + NSET}. The stored previous value is compared with the current new value to decide whether a \ttt{PDFSET} call is needed to switch to another set. \iteme{MINT(101), MINT(102) :} is normally 1, but is 4 when a resolved photon (appearing on side 1 or 2) can be represented by either of the four vector mesons $\rho^0$, $\omega$, $\phi$ and $\mathrm{J}/\psi$. \iteme{MINT(103), MINT(104) :} KF flavour code for the two incoming particles, i.e. the same as \ttt{MINT(11)} and \ttt{MINT(12)}. The exception is when a resolved photon is represented by a vector meson (a $\rho^0$, $\omega$, $\phi$ or $\mathrm{J}/\psi$). Then the code of the vector meson is given. \iteme{MINT(105) :} is either \ttt{MINT(103)} or \ttt{MINT(104)}, depending on which side of the event currently is being studied. \iteme{MINT(107), MINT(108) :} if either or both of the two incoming particles is a photon, then the respective value gives the nature assumed for that photon. The code follows the one used for \ttt{MSTP(14)}: \begin{subentry} \iteme{= 0 :} direct photon. \iteme{= 1 :} resolved photon. \iteme{= 2 :} VMD-like photon. \iteme{= 3 :} anomalous photon. \end{subentry} \iteme{MINT(109) :} is either \ttt{MINT(107)} or \ttt{MINT(108)}, depending on which side of the event currently is being studied. \iteme{MINT(111) :} the frame given in \ttt{PYINIT} call, 0--5 for \ttt{'NONE'}, \ttt{'CMS'}, \ttt{'FIXT'}, \ttt{'USER'}, \ttt{'FOUR'} and \ttt{'FIVE'}, respectively. \iteme{MINT(121) :} number of separate event classes to initialize and mix. \begin{subentry} \iteme{= 1 :} the normal value. \iteme{= 3 :} for a $\gamma\mathrm{p}$ interaction when \ttt{MSTP(14)=10}. \iteme{= 6 :} for a $\gamma\gamma$ interaction when \ttt{MSTP(14)=10}. \end{subentry} \iteme{MINT(122) :} event class used in current event for $\gamma\mathrm{p}$ or $\gamma\gamma$ events generated with the \ttt{MSTP(14)=10} option; code as described for \ttt{MSTI(9)}. \iteme{MINT(123) :} event class used in the current event, with the same list of possibilities as for \ttt{MSTP(14)}, except that options 1, 4 or 10 do not appear. Apart from a different coding, this is exactly the same information as is available in \ttt{MINT(122)}. \boxsep \iteme{VINT(1) :}\label{p:VINT} $E_{\mrm{cm}}$, c.m. energy. \iteme{VINT(2) :} $s$ ($=E_{\mrm{cm}}^2$) squared mass of complete system. \iteme{VINT(3) :} mass of beam particle. \iteme{VINT(4) :} mass of target particle. \iteme{VINT(5) :} momentum of beam (and target) particle in c.m. frame. \iteme{VINT(6) - VINT(10) :} $\theta$, $\varphi$ and \mbox{{\boldmath $\beta$}} for rotation and boost from c.m. frame to user-specified frame. \iteme{VINT(11) :} $\tau_{\mathrm{min}}$. \iteme{VINT(12) :} $y_{\mathrm{min}}$. \iteme{VINT(13) :} $\cos\hat{\theta}_{\mathrm{min}}$ for $\cos\hat{\theta} \leq 0$. \iteme{VINT(14) :} $\cos\hat{\theta}_{\mathrm{min}}$ for $\cos\hat{\theta} \geq 0$. \iteme{VINT(15) :} $x^2_{\perp \mathrm{min}}$. \iteme{VINT(16) :} $\tau'_{\mathrm{min}}$. \iteme{VINT(21) :} $\tau$. \iteme{VINT(22) :} $y$. \iteme{VINT(23) :} $\cos\hat{\theta}$. \iteme{VINT(24) :} $\varphi$ (azimuthal angle). \iteme{VINT(25) :} $x_{\perp}^2$. \iteme{VINT(26) :} $\tau'$. \iteme{VINT(31) :} $\tau_{\mathrm{max}}$. \iteme{VINT(32) :} $y_{\mathrm{max}}$. \iteme{VINT(33) :} $\cos\hat{\theta}_{\mathrm{max}}$ for $\cos\hat{\theta} \leq 0$. \iteme{VINT(34) :} $\cos\hat{\theta}_{\mathrm{max}}$ for $\cos\hat{\theta} \geq 0$. \iteme{VINT(35) :} $x^2_{\perp \mathrm{max}}$. \iteme{VINT(36) :} $\tau'_{\mathrm{max}}$. \iteme{VINT(41), VINT(42) :} the momentum fractions $x$ taken by the partons at the hard interaction, as used e.g. in the parton-distribution functions. \iteme{VINT(43) :} $\hat{m} = \sqrt{\hat{s}}$, mass of hard-scattering subsystem. \iteme{VINT(44) :} $\hat{s}$ of the hard subprocess ($2 \to 2$ or $2 \to 1$). \iteme{VINT(45) :} $\hat{t}$ of the hard subprocess ($2 \to 2$ or $2 \to 1 \to 2$). \iteme{VINT(46) :} $\hat{u}$ of the hard subprocess ($2 \to 2$ or $2 \to 1 \to 2$). \iteme{VINT(47) :} $\hat{p}_{\perp}$ of the hard subprocess ($2 \to 2$ or $2 \to 1 \to 2$), i.e. transverse momentum evaluated in the rest frame of the scattering. \iteme{VINT(48) :} $\hat{p}_{\perp}^2$ of the hard subprocess; see \ttt{VINT(47)}. \iteme{VINT(49) :} $\hat{m}'$, the mass of the complete three- or four-body final state in $2 \to 3$ or $2 \to 4$ processes. \iteme{VINT(50) :} $\hat{s}' = \hat{m}'^2$; see \ttt{VINT(49)}. \iteme{VINT(51) :} $Q$ of the hard subprocess. The exact definition is process-dependent, see \ttt{MSTP(32)}. \iteme{VINT(52) :} $Q^2$ of the hard subprocess; see \ttt{VINT(51)}. \iteme{VINT(53) :} $Q$ of the outer hard-scattering subprocess. Agrees with \ttt{VINT(51)} for a $2 \to 1$ or $2 \to 2$ process. For a $2 \to 3$ or $2 \to 4$ $\mathrm{W}/\mathrm{Z}$ fusion process, it is set by the $\mathrm{W}/\mathrm{Z}$ mass scale, and for subprocesses 121 and 122 by the heavy-quark mass. \iteme{VINT(54) :} $Q^2$ of the outer hard-scattering subprocess; see \ttt{VINT(53)}. \iteme{VINT(55) :} $Q$ scale used as maximum virtuality in parton showers. Is equal to \ttt{VINT(53)}, except for deep-inelastic-scattering processes when \ttt{MSTP(22)}$ > 0$. \iteme{VINT(56) :} $Q^2$ scale in parton showers; see \ttt{VINT(55)}. \iteme{VINT(57) :} $\alpha_{\mathrm{em}}$ value of hard process. \iteme{VINT(58) :} $\alpha_{\mathrm{s}}$ value of hard process. \iteme{VINT(59) :} $\sin\hat{\theta}$ (cf. \ttt{VINT(23)}); used for improved numerical precision in elastic and diffractive scattering. \iteme{VINT(61), VINT(62) :} nominal $m^2$ values, i.e. without initial-state radiation effects, for the two partons entering the hard interaction. \iteme{VINT(63), VINT(64) :} nominal $m^2$ values, i.e. without final-state radiation effects, for the two (or one) partons/particles leaving the hard interaction. \iteme{VINT(65) :} $\hat{p}_{\mrm{init}}$, i.e. common nominal absolute momentum of the two partons entering the hard interaction, in their rest frame. \iteme{VINT(66) :} $\hat{p}_{\mrm{fin}}$, i.e. common nominal absolute momentum of the two partons leaving the hard interaction, in their rest frame. \iteme{VINT(67), VINT(68) :} mass of beam and target particle, as \ttt{VINT(3)} and \ttt{VINT(4)}, except that an incoming $\gamma$ is assigned the $\rho^0$, $\omega$ or $\phi$ mass. Used for elastic scattering $\gamma \mathrm{p} \to \rho^0 \mathrm{p}$ and other similar processes. \iteme{VINT(71) :} $p_{\perp\mathrm{min}}$ of process, i.e. \ttt{CKIN(3)} or \ttt{CKIN(5)}, depending on which is larger, and whether the process is singular in $p_{\perp} \to 0$ or not. \iteme{VINT(73) :} $\tau = m^2/s$ value of first resonance, if any; see \ttt{MINT(73)}. \iteme{VINT(74) :} $m \Gamma/s$ value of first resonance, if any; see \ttt{MINT(73)}. \iteme{VINT(75) :} $\tau = m^2/s$ value of second resonance, if any; see \ttt{MINT(74)}. \iteme{VINT(76) :} $m \Gamma/s$ value of second resonance, if any; see \ttt{MINT(74)}. \iteme{VINT(80) :} correction factor (evaluated in \ttt{PYOFSH}) for the cross section of resonances produced in $2 \to 2$ processes, if only some mass range of the full Breit--Wigner shape is allowed by user-set mass cuts (\ttt{CKIN(2)}, \ttt{CKIN(45) - CKIN(48)}). \iteme{VINT(81) - VINT(84) :} the $\cos\theta$ and $\varphi$ variables of a true $2 \to 3$ process, where one product is a resonance, effectively giving $2 \to 4$. The first two are $\cos\theta$ and $\varphi$ for the resonance decay, the other two ditto for the effective system formed by the other two particles. \iteme{VINT(85), VINT(86) :} transverse momenta in a true $2 \to 3$ process; one is stored in \ttt{VINT(47)} (that of the $\mathrm{Z}^0$ in $\mathrm{g} \mathrm{g} \to \mathrm{Z}^0 \mathrm{Q} \overline{\mathrm{Q}}$), while the smaller of the other two is stored in \ttt{VINT(85)} and the larger in \ttt{VINT(86)}. \iteme{VINT(91), VINT(92) :} gives a dimensionless suppression factor, to take into account reduction in cross section due to the allowed channels for a $\mathrm{W}^+\mathrm{W}^+$ or $\mathrm{W}^-\mathrm{W}^-$ pair, respectively, in the same sense as \ttt{WIDS(24,1)} gives it for a $\mathrm{W}^+\mathrm{W}^-$ pair. \iteme{VINT(98) :} is sum of \ttt{VINT(100)} values for current run. \iteme{VINT(99) :} is weight \ttt{WTXS} returned from \ttt{PYEVWT} call when \ttt{MSTP(142)}$\geq 1$, otherwise is 1. \iteme{VINT(100) :} is compensating weight \ttt{1./WTXS} that should be associated with events when \ttt{MSTP(142)=1}, otherwise is 1. \iteme{VINT(108) :} ratio of maximum differential cross section observed to maximum differential cross section assumed for the generation; cf. \ttt{MSTP(123)}. \iteme{VINT(109) :} ratio of minimal (negative!) cross section observed to maximum differential cross section assumed for the generation; could only become negative if cross sections are incorrectly included. \iteme{VINT(111) - VINT(116) :} for \ttt{MINT(61)=1} gives kinematical factors for the different pieces contributing to $\gamma^* / \Z^0$ or $\gamma^*/\mathrm{Z}^0/\mathrm{Z}'^0$ production, for \ttt{MINT(61)=2} gives sum of final-state weights for the same; coefficients are given in the order pure $\gamma^*$, $\gamma^*$--$\mathrm{Z}^0$ interference, $\gamma^*$--$\mathrm{Z}'^0$ interference, pure $\mathrm{Z}^0$, $\mathrm{Z}^0$--$\mathrm{Z}'^0$ interference and pure $\mathrm{Z}'^0$. \iteme{VINT(117) :} width of $\mathrm{Z}^0$; needed in $\gamma^*/\mathrm{Z}^0/\mathrm{Z}'^0$ production. \iteme{VINT(131) :} total cross section (in mb) for subprocesses allowed in the pile-up events scenario according to the \ttt{MSTP(132)} value. \iteme{VINT(132) :} $\br{n} = $\ttt{VINT(131)}$\times$\ttt{PARP(131)}, cf. \ttt{PARI(91)}. \iteme{VINT(133) :} $\langle n \rangle = \sum_i i \, {\cal P}_i / \sum_i {\cal P}_i$ as actually simulated, i.e. $1 \leq i \leq 200$ (or smaller), see \ttt{PARI(92)}. \iteme{VINT(134) :} number related to probability to have event in beam--beam crossing; is $\exp(-\br{n}) \sum_i \br{n}^i/i!$ for \ttt{MSTP(133)=1} and $\exp(-\br{n}) \sum_i \br{n}^i/(i-1)!$ for \ttt{MSTP(133)=2}, cf. \ttt{PARI(93)}. \iteme{VINT(138) :} size of the threshold factor (enhancement or suppression) in the latest event with heavy-flavour production; see \ttt{MSTP(35)}. \iteme{VINT(141), VINT(142) :} $x$ values for the parton-shower initiators of the hardest interaction; used to find what is left for multiple interactions. \iteme{VINT(143), VINT(144) :} $1 - \sum_i x_i$ for all scatterings; used for rescaling each new $x$-value in the multiple-interaction parton-distribution-function evaluation. \iteme{VINT(145) :} estimate of total parton--parton cross section for multiple interactions; used for \ttt{MSTP(82)}$\geq 2$. \iteme{VINT(146) :} common correction factor $f_c$ in the multiple-interaction probability; used for \ttt{MSTP(82)}$\geq 2$ (part of $e(b)$, see eq.~(\ref{mi:ebenh})). \iteme{VINT(147) :} average hadronic matter overlap; used for \ttt{MSTP(82)}$\geq 2$ (needed in evaluation of $e(b)$, see eq.~(\ref{mi:ebenh})). \iteme{VINT(148) :} enhancement factor for current event in the multiple-interaction probability, defined as the actual overlap divided by the average one; used for \ttt{MSTP(82)}$\geq 2$ (is $e(b)$ of eq.~(\ref{mi:ebenh})). \iteme{VINT(149) :} $x_{\perp}^2$ cut-off or turn-off for multiple interactions. For \ttt{MSTP(82)}$\leq 1$ it is $4 p_{\perp\mathrm{min}}^2/s$, for \ttt{MSTP(82)}$\geq 2$ it is $4 p_{\perp 0}^2/s$. \iteme{VINT(150) :} probability to keep the given event in the multiple-interaction scenario, as given by the `Sudakov' form factor. \iteme{VINT(151), VINT(152) :} sum of $x$ values for all the multiple-interaction partons. \iteme{VINT(153) :} current differential cross section value obtained from \ttt{PYSIGH}; used in multiple interactions only. \iteme{VINT(155), VINT(156) :} the $x$ value of a photon that branches into quarks or gluons, i.e. $x$ at interface between initial-state QED and QCD cascades. \iteme{VINT(157), VINT(158) :} the primordial $k_{\perp}$ values selected in the two beam remnants. \iteme{VINT(159), VINT(160) :} the $\chi$ values selected for beam remnants that are split into two objects, describing how the energy is shared (see \ttt{MSTP(92)} and \ttt{MSTP(94)}); is 0 if no splitting is needed. \iteme{VINT(161) - VINT(200) :} sum of Cabibbo--Kobayashi--Maskawa squared matrix elements that a given flavour is allowed to couple to. Results are stored in format \ttt{VINT(180+KF)} for quark and lepton flavours and antiflavours (which need not be the same; see \ttt{MDME(IDC,2)}). For leptons, these factors are normally unity. \iteme{VINT(201) - VINT(220) :} additional variables needed in phase-space selection for $2 \to 3$ processes with \ttt{ISET(ISUB)=5}. Below indices 1, 2 and 3 refer to scattered partons 1, 2 and 3, except that the $q$ four-momentum variables are $q_1 + q_2 \to q_1' q_2' q_3'$. All kinematical variables refer to the internal kinematics of the 3-body final state --- the kinematics of the system as a whole is described by $\tau'$ and $y$, and the mass distribution of particle 3 (a resonance) by $\tau$. \begin{subentry} \iteme{VINT(201) :} $m_1$. \iteme{VINT(202) :} $p_{\perp 1}^2$. \iteme{VINT(203) :} $\varphi_1$. \iteme{VINT(204) :} $M_1$ (mass of propagator particle). \iteme{VINT(205) :} weight for the $p_{\perp 1}^2$ choice. \iteme{VINT(206) :} $m_2$. \iteme{VINT(207) :} $p_{\perp 2}^2$. \iteme{VINT(208) :} $\varphi_2$. \iteme{VINT(209) :} $M_2$ (mass of propagator particle). \iteme{VINT(210) :} weight for the $p_{\perp 2}^2$ choice. \iteme{VINT(211) :} $y_3$. \iteme{VINT(212) :} $y_{3 \mathrm{max}}$. \iteme{VINT(213) :} $\epsilon = \pm 1$; choice between two mirror solutions $1 \leftrightarrow 2$. \iteme{VINT(214) :} weight associated to $\epsilon$-choice. \iteme{VINT(215) :} $t_1 = (q_1 - q_1')^2$. \iteme{VINT(216) :} $t_2 = (q_2 - q_2')^2$. \iteme{VINT(217) :} $q_1 q_2'$ four-product. \iteme{VINT(218) :} $q_2 q_1'$ four-product. \iteme{VINT(219) :} $q_1' q_2'$ four-product. \iteme{VINT(220) :} $\sqrt{(m_{\perp 12}^2 - m_{\perp 1}^2 - m_{\perp 2}^2)^2 - 4 m_{\perp 1}^2 m_{\perp 2}^2}$, where $m_{\perp 12}$ is the transverse mass of the $q'_1 q'_2$ system. \end{subentry} \iteme{VINT(221) - VINT(225) :} $\theta$, $\varphi$ and \mbox{{\boldmath $\beta$}} for rotation and boost from c.m. frame to hadronic c.m. frame of a lepton--hadron event. \iteme{VINT(231) :} $Q^2_{\mathrm{min}}$ scale for current parton-distribution function set. \iteme{VINT(281) :} for resolved photon events, it gives the ratio between the total $\gamma X$ cross section and the total $\pi^0 X$ cross section, where $X$ represents the target particle. \iteme{VINT(283), VINT(284) :} virtuality scale at which an anomalous photon on the beam or target side of the event is being resolved. More precisely, it gives the $p_{\perp}^2$ of the $\gamma \to \mathrm{q}\overline{\mathrm{q}}$ vertex. \iteme{VINT(285) :} the \ttt{CKIN(3)} value provided by the user at initialization; subsequently \ttt{CKIN(3)} may be overwritten (for \ttt{MSTP(14)=10}) but \ttt{VINT(285)} stays. \iteme{VINT(289) :} squared c.m. energy found in \ttt{PYINIT} call. \iteme{VINT(290) :} the \ttt{WIN} argument of a \ttt{PYINIT} call. \iteme{VINT(291) - VINT(300) :} the two five-vectors of the two incoming particles, as reconstructed in \ttt{PYINKI}. These may vary from one event to the next. \end{entry} \drawbox{COMMON/PYINT2/ISET(200),KFPR(200,2),COEF(200,20),ICOL(40,4,2)}% \label{p:PYINT2} \begin{entry} \itemc{Purpose:} to store information necessary for efficient generation of the different subprocesses, specifically type of generation scheme and coefficients of the Jacobian. Also to store allowed colour-flow configurations. These variables must not be changed by you. \iteme{ISET(ISUB) :}\label{p:ISET} gives the type of kinematical-variable selection scheme used for subprocess ISUB. \begin{subentry} \iteme{= 0 :} elastic, diffractive and low-$p_{\perp}$ processes. \iteme{= 1 :} $2 \to 1$ processes (irrespective of subsequent decays). \iteme{= 2 :} $2 \to 2$ processes (i.e. the bulk of processes). \iteme{= 3 :} $2 \to 3$ processes (like $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \mathrm{q}''' \H^0$). \iteme{= 4 :} $2 \to 4$ processes (like $\mathrm{q} \mathrm{q}' \to \mathrm{q}'' \mathrm{q}''' \mathrm{W}^+ \mathrm{W}^-$). \iteme{= 5 :} `true' $2 \to 3$ processes, one method. \iteme{= 6 :} `true' $2 \to 3$ processes, another method; currently only $\mathrm{g} \mathrm{g} \to \mathrm{Z}^0 \mathrm{Q} \overline{\mathrm{Q}}$. \iteme{= 9 :} $2 \to 2$ in multiple interactions ($p_{\perp}$ as kinematics variable). \iteme{= 11 :} a user-defined process. \iteme{= -1 :} legitimate process which has not yet been implemented. \iteme{= -2 :} ISUB is an undefined process code. \end{subentry} \iteme{KFPR(ISUB,J) :}\label{p:KFPR} give the KF flavour codes for the products produced in subprocess ISUB. If there is only one product, the \ttt{J=2} position is left blank. Also, quarks and leptons assumed massless in the matrix elements are denoted by 0. The main application is thus to identify resonances produced ($\mathrm{Z}^0$, $\mathrm{W}^{\pm}$, $\H^0$, etc.). \iteme{COEF(ISUB,J) :}\label{p:COEF} factors used in the Jacobians in order to speed up the selection of kinematical variables. More precisely, the shape of the cross section is given as the sum of terms with different behaviour, where the integral over the allowed phase space is unity for each term. \ttt{COEF} gives the relative strength of these terms, normalized so that the sum of coefficients for each variable used is unity. Note that which coefficients are indeed used is process-dependent. \begin{subentry} \iteme{ISUB :} standard subprocess code. \iteme{J = 1 :} $\tau$ selected according $1/\tau$. \iteme{J = 2 :} $\tau$ selected according to $1/\tau^2$. \iteme{J = 3 :} $\tau$ selected according to $1/(\tau(\tau+\tau_R))$, where $\tau_R = m_R^2/s$ is $\tau$ value of resonance; only used for resonance production. \iteme{J = 4 :} $\tau$ selected according to Breit--Wigner of form $1/((\tau-\tau_R)^2+\gamma_R^2)$, where $\tau_R = m_R^2/s$ is $\tau$ value of resonance and $\gamma_R = m_R \Gamma_R/s$ is its scaled mass times width; only used for resonance production. \iteme{J = 5 :} $\tau$ selected according to $1/(\tau(\tau+\tau_{R'}))$, where $\tau_{R'} = m_{R'}^2/s$ is $\tau$ value of second resonance; only used for simultaneous production of two resonances. \iteme{J = 6 :} $\tau$ selected according to second Breit--Wigner of form $1/((\tau-\tau_{R'})^2+\gamma_{R'}^2)$, where $\tau_{R'} = m_{R'}^2/s$ is $\tau$ value of second resonance and $\gamma_{R'} = m_{R'} \Gamma_{R'}/s$ is its scaled mass times width; is used only for simultaneous production of two resonances, like $\gamma^*/\mathrm{Z}^0/\mathrm{Z}'^0$. \iteme{J = 7 :} $\tau$ selected according to $1/(1-\tau)$; only used when both parton distributions are peaked at $x = 1$. \iteme{J = 8 :} $y$ selected according to $y - y_{\mathrm{min}}$. \iteme{J = 9 :} $y$ selected according to $y_{\mathrm{max}} - y$. \iteme{J = 10 :} $y$ selected according to $1/\cosh(y)$. \iteme{J = 11 :} $y$ selected according to $1/(1-\exp(y-y_{\mathrm{max}}))$; only used when beam parton distribution is peaked close to $x = 1$. \iteme{J = 12 :} $y$ selected according to $1/(1-\exp(y_{\mathrm{min}}-y))$; only used when target parton distribution is peaked close to $x = 1$. \iteme{J = 13 :} $z = \cos\hat{\theta}$ selected evenly between limits. \iteme{J = 14 :} $z = \cos\hat{\theta}$ selected according to $1/(a-z)$, where $a = 1 + 2 m_3^2 m_4^2/\hat{s}^2$, $m_3$ and $m_4$ being the masses of the two final-state particles. \iteme{J = 15 :} $z = \cos\hat{\theta}$ selected according to $1/(a+z)$, with $a$ as above. \iteme{J = 16 :} $z = \cos\hat{\theta}$ selected according to $1/(a-z)^2$, with $a$ as above. \iteme{J = 17 :} $z = \cos\hat{\theta}$ selected according to $1/(a+z)^2$, with $a$ as above. \iteme{J = 18 :} $\tau'$ selected according to $1/\tau'$. \iteme{J = 19 :} $\tau'$ selected according to $(1 - \tau/\tau')^3/\tau'^2$. \iteme{J = 20 :} $\tau'$ selected according to $1/(1-\tau')$; only used when both parton distributions are peaked close to $x = 1$. \end{subentry} \iteme{ICOL :}\label{p:ICOL} contains information on different colour-flow topologies in hard $2 \to 2$ processes. \end{entry} \drawbox{COMMON/PYINT3/XSFX(2,-40:40),ISIG(1000,3),SIGH(1000)}% \label{p:PYINT3} \begin{entry} \itemc{Purpose:} to store information on parton distributions, subprocess cross sections and different final-state relative weights. These variables must not be changed by you. \iteme{XSFX :}\label{p:XSFX} current values of parton-distribution functions $xf(x)$ on beam and target side. \iteme{ISIG(ICHN,1) :}\label{p:ISIG} incoming parton/particle on the beam side to the hard interaction for allowed channel number \ttt{ICHN}. The number of channels filled with relevant information is given by \ttt{NCHN}, one of the arguments returned in a \ttt{PYSIGH} call. Thus only $1 \leq $\ttt{ICHN}$ \leq $\ttt{NCHN} is filled with relevant information. \iteme{ISIG(ICHN,2) :} incoming parton/particle on the target side to the hard interaction for allowed channel number \ttt{ICHN}. See also comment above. \iteme{ISIG(ICHN,3) :} colour-flow type for allowed channel number \ttt{ICHN}; see \ttt{MSTI(2)} list. See also above comment. For `subprocess' 96 uniquely, \ttt{ISIG(ICHN,3)} is also used to translate information on what is the correct subprocess number (11, 12, 13, 28, 53 or 68); this is used for reassigning subprocess 96 to either of these. \iteme{SIGH(ICHN) :}\label{p:SIGH} evaluated differential cross section for allowed channel number \ttt{ICHN}, i.e. matrix-element value times parton distributions, for current kinematical setup (in addition, Jacobian factors are included in the figures, as used to speed up generation). See also comment for \ttt{ISIG(ICHN,1)}. \end{entry} \drawbox{COMMON/PYINT4/WIDP(21:40,0:40),WIDE(21:40,0:40),WIDS(21:40,3)}% \label{p:PYINT4} \begin{entry} \itemc{Purpose:} to store partial and effective decay widths for the different resonances. These variables must not be changed by you. \iteme{WIDP(KF,J) :}\label{p:WIDP} gives partial decay widths of resonances into different channels (in GeV), given that all physically allowed final states are included. \begin{subentry} \iteme{KF :} standard KF code for resonance considered. When top is treated like a resonance (see \ttt{MSTP(48)}) it is stored in position 26. When the fourth generation fermions $\mathrm{l}$, $\mathrm{h}$, $\chi^-$ and $\nu_{\chi}$ are treated like resonances (see \ttt{MSTP(49)}) they are stored in positions 27, 28, 29 and 30, respectively. \iteme{J :} enumerates the different decay channels possible for resonance KF, as stored in the {\tsc{Jetset}} \ttt{LUDAT3} common block, with the first channel in \ttt{J=1}, etc. \end{subentry} \iteme{WIDE(KF,J) :}\label{p:WIDE} gives effective decay widths of resonances into different channels (in GeV), given the decay modes actually left open in the current run. The on/off status of decay modes is set by the \ttt{MDME} switches in {\tsc{Jetset}}; see section \ref{ss:parapartdat}. \begin{subentry} \iteme{KF :} standard KF code for resonance considered. For comment about top and fourth generation see \ttt{WIDP} above. \iteme{J :} enumerates the different decay channels possible for resonance KF, as stored in the {\tsc{Jetset}} \ttt{LUDAT3} common block, with the first channel in \ttt{J=1}, etc. \end{subentry} \iteme{WIDS(KF,J) :}\label{p:WIDS} gives a dimensionless suppression factor, which is defined as the ratio of the total width of channels switched on to the total width of all possible channels (replace width by squared width for a pair of resonances). The on/off status of channels is set by the \ttt{MDME} switches in {\tsc{Jetset}}; see section \ref{ss:parapartdat}. The information in \ttt{WIDS} is used e.g. in cross-section calculations. \begin{subentry} \iteme{KF :} standard KF code for resonance considered. For comment about top and fourth generation see \ttt{WIDP} above. \iteme{J = 1 :} suppression when a pair of resonances of type KF are produced together. When an antiparticle exists, the particle--antiparticle pair (such as $\mathrm{W}^+ \mathrm{W}^-$) is the relevant combination, else the particle--particle one (such as $\mathrm{Z}^0 \mathrm{Z}^0$). \iteme{J = 2 :} suppression for a particle of type KF when produced on its own, or together with a particle of another type. \iteme{J = 3 :} suppression for an antiparticle of type KF when produced on its own, or together with a particle of another type. \end{subentry} \end{entry} \drawbox{COMMON/PYINT5/NGEN(0:200,3),XSEC(0:200,3)}\label{p:PYINT5} \begin{entry} \itemc{Purpose:} to store information necessary for cross-section calculation and differential cross-section maximum violation. These variables must not be changed by you. \iteme{NGEN(ISUB,1) :}\label{p:NGEN} gives the number of times that the differential cross section (times Jacobian factors) has been evaluated for subprocess ISUB, with \ttt{NGEN(0,1)} the sum of these. \iteme{NGEN(ISUB,2) :} gives the number of times that a kinematical setup for subprocess ISUB is accepted in the generation procedure, with \ttt{NGEN(0,2)} the sum of these. \iteme{NGEN(ISUB,3) :} gives the number of times an event of subprocess type ISUB is generated, with \ttt{NGEN(0,3)} the sum of these. Usually \ttt{NGEN(ISUB,3) = NGEN(ISUB,2)}, i.e. an accepted kinematical configuration can normally be used to produce an event. \iteme{XSEC(ISUB,1) :}\label{p:XSEC} estimated maximum differential cross section (times the Jacobian factors used to speed up the generation process) for the different subprocesses in use, with \ttt{XSEC(0,1)} the sum of these (except low-$p_{\perp}$, i.e. ISUB = 95). \iteme{XSEC(ISUB,2) :} gives the sum of differential cross sections (times Jacobian factors) for the \ttt{NGEN(ISUB,1)} phase-space points evaluated so far. \iteme{XSEC(ISUB,3) :} gives the estimated integrated cross section for subprocess ISUB, based on the statistics accumulated so far, with \ttt{XSEC(0,3)} the estimated total cross section for all subprocesses included (all in mb). This is exactly the information obtainable by a \ttt{PYSTAT(1)} call. \end{entry} \drawboxtwo{COMMON/PYINT6/PROC(0:200)}{CHARACTER PROC*28}% \label{p:PYINT6} \begin{entry} \itemc{Purpose:} to store character strings for the different possible subprocesses; used when printing tables. \iteme{PROC(ISUB) :}\label{p:PROC} name for the different subprocesses, according to ISUB code. \ttt{PROC(0)} denotes all processes. \end{entry} \drawbox{COMMON/PYINT7/SIGT(0:6,0:6,0:5)}\label{p:PYINT7} \begin{entry} \itemc{Purpose:} to store information on total, elastic and diffractive cross sections. These variables should only be set by you for the option \ttt{MSTP(31)=0}; else they should not be touched. All numbers are given in mb. \iteme{SIGT(I1,I2,J) :}\label{p:SIGT} the cross section, both total and subdivided by class (elastic, diffractive etc.). For a photon to be considered as a VMD meson the cross sections are additionally split into the contributions from the various meson states. \begin{subentry} \iteme{I1, I2 :} allowed states for the incoming particle on side 1 and 2, respectively. \begin{subentry} \iteme{= 0 :} sum of all allowed states. Except for a photon to be considered as a VMD meson this is the only nonvanishing entry. \iteme{= 1 :} the contribution from the $\rho^0$ VMD state. \iteme{= 2 :} the contribution from the $\omega$ VMD state. \iteme{= 3 :} the contribution from the $\phi$ VMD state. \iteme{= 4 :} the contribution from the $\mathrm{J}/\psi$ VMD state. \iteme{= 5, 6 :} reserved for future use. \end{subentry} \iteme{J :} the total and partial cross sections. \begin{subentry} \iteme{= 0 :} the total cross section. \iteme{= 1 :} the elastic cross section. \iteme{= 2 :} the single diffractive cross section $AB \to XB$. \iteme{= 3 :} the single diffractive cross section $AB \to AX$. \iteme{= 4 :} the double diffractive cross section. \iteme{= 5 :} the inelastic, non-diffractive cross section. \end{subentry} \itemc{Warning:} If you set these values yourself, it is important that they are internally consistent, since this is not explicitly checked by the program. Thus the contributions \ttt{J=}1--5 should add up to the \ttt{J=}0 one and, for VMD photons, the contributions \ttt{I=}1--4 should add up to the \ttt{I=}0 one. \end{subentry} \end{entry} \drawboxtwo{~COMMON/PYINT8/XPVMD(-6:6),XPANL(-6:6),XPANH(-6:6),% XPBEH(-6:6),}{\&XPDIR(-6:6)}\label{p:PYINT8} \begin{entry} \itemc{Purpose:} to store the various components of the photon parton distributions when the \ttt{PYGGAM} routine is called. \iteme{XPVMD(KFL) :}\label{p:XPVMD} gives distributions of the VMD part ($\rho^0$, $\omega$ and $\phi$). \iteme{XPANL(KFL) :}\label{p:XPANL} gives distributions of the anomalous part of light quarks ($\d$, $\u$ and $\mathrm{s}$). \iteme{XPANH(KFL) :}\label{p:XPANH} gives distributions of the anomalous part of heavy quarks ($\c$ and $\b$). \iteme{XPBEH(KFL) :}\label{p:XPBEH} gives Bethe-Heitler distributions of heavy quarks ($\c$ and $\b$). This provides an alternative to \ttt{XPANH}, i.e. both should not be used at the same time. \iteme{XPDIR(KFL) :}\label{p:XPDIR} gives direct correction to the production of light quarks ($\d$, $\u$ and $\mathrm{s}$). This term is nonvanishing only in the {$\overline{\mrm{MS}}$} scheme, and is applicable for $F_2^{\gamma}$ rather than for the parton distributions themselves. \end{entry} \boxsep Finally, in addition a number of routines and common blocks with names beginning with \ttt{RK} come with the program. These contain the matrix-element evaluation for the process $\mathrm{g} \mathrm{g} \to \mathrm{Z} \mathrm{q} \overline{\mathrm{q}}$, based on a program of Ronald Kleiss, with only minor modifications. \subsection{Examples} The program is built as a slave system, i.e. you supply the main program, which calls on the {\tsc{Pythia}} and {\tsc{Jetset}} routines to perform specific tasks and then resumes control. A typical program for the analysis of collider events at 630 GeV c.m. energy with a minimum $p_{\perp}$ of 10 GeV/c at the hard scattering (because of initial-state radiation, fragmentation effects, etc., the actual $p_{\perp}$ cut-off will be smeared around this value) might look like \begin{verbatim} COMMON/LUJETS/N,K(4000,5),P(4000,5),V(4000,5) COMMON/PYSUBS/MSEL,MSUB(200),KFIN(2,-40:40),CKIN(200) COMMON/PYPARS/MSTP(200),PARP(200),MSTI(200),PARI(200) ... ! set all common block variables that ... ! did not have desired default values CKIN(3)=10. ! lower p_T cut-off CALL PYINIT('CMS','p','pbar',630.) ! initialize ... ! initialize analysis statistics DO 100 IEVENT=1,1000 ! loop over events CALL PYEVNT ! generate event IF(IEVENT.EQ.1) CALL LULIST(1) ! list first event ... ! insert desired analysis chain for ... ! each event 100 CONTINUE CALL PYSTAT(1) ! print cross sections ... ! user output END \end{verbatim} \clearpage \section{Initial- and Final-State Radiation} \label{s:showinfi} Starting from the hard interaction, initial- and final-state radiation corrections may be added. This is normally done by making use of the parton-shower language --- only for the $\e^+\e^- \to \mathrm{q} \overline{\mathrm{q}}$ process does {\tsc{Jetset}} offer a matrix-element option (described in section \ref{ss:eematrix}). The algorithms used to generate initial- and final-state showers are rather different, and are therefore described separately below, starting with the conceptually easier final-state one. Before that, some common elements are introduced. The main reference for final-state showers is ref. \cite{Ben87a} and for initial-state ones ref. \cite{Sjo85}. \subsection{Shower Evolution} In the leading log picture, a shower may be viewed as a sequence of $1 \to 2$ branchings $a \to bc$. Here $a$ is called the mother and $b$ and $c$ the two daughters. Each daughter is free to branch in its turn, so that a tree-like stucture can evolve. We will use the work `parton' for all the objects $a$, $b$ and $c$ involved in the branching process, i.e. not only for quarks and gluons but also for leptons and photons. The branchings included in the program are $\mathrm{q} \to \mathrm{q} \mathrm{g}$, $\mathrm{g} \to \mathrm{g} \mathrm{g}$, $\mathrm{g} \to \mathrm{q} \overline{\mathrm{q}}$, $\mathrm{q} \to \mathrm{q} \gamma$ and $\ell \to \ell \gamma$. Photon branchings, i.e. $\gamma \to \mathrm{q} \overline{\mathrm{q}}$ and $\gamma \to \ell \br{\ell}$, have not been included so far, since they are reasonably rare and since no urgent need for them has been perceived. \subsubsection{The evolution equations} In the shower formulation, the kinematics of each branching is given in terms of two variables, $Q^2$ and $z$. Slightly different interpretations may be given to these variables, and indeed this is one main area where the various programs on the market differ. $Q^2$ has dimensions of squared mass, and is related to the mass or transverse momentum scale of the branching. $z$ gives the sharing of the $a$ energy and momentum between the two daughters, with parton $b$ taking a fraction $z$ and parton $c$ a fraction $1-z$. To specify the kinematics, an azimuthal angle $\varphi$ of the $b$ around the $a$ direction is needed in addition; normally $\varphi$ is chosen to be isotropically distributed, although options for non-isotropic distributions exist. The probability for a parton to branch is given by the evolution equations (also called DGLAP or Altarelli--Parisi \cite{Gri72,Alt77}). It is convenient to introduce \begin{equation} t = \ln(Q^2/\Lambda^2) ~~~ \Rightarrow ~~~ \d t = \d \ln(Q^2) = \frac{\d Q^2}{Q^2} ~, \end{equation} where $\Lambda$ is the QCD $\Lambda$ scale in $\alpha_{\mathrm{s}}$. Of course, this choice is more directed towards the QCD parts of the shower, but it can be used just as well for the QED ones. In terms of the two variables $t$ and $z$, the differential probability $\d {\cal P}$ for parton $a$ to branch is now \begin{equation} \d {\cal P}_a = \sum_{b,c} \frac{\alpha_{abc}}{2 \pi} \, P_{a \to bc}(z) \, \d t \, \d z ~. \label{sh:Patobc} \end{equation} Here the sum is supposed to run over all allowed branchings, for a quark $\mathrm{q} \to \mathrm{q} \mathrm{g}$ and $\mathrm{q} \to \mathrm{q} \gamma$, and so on. The $\alpha_{abc}$ factor is $\alpha_{\mathrm{em}}$ for QED branchings and $\alpha_{\mathrm{s}}$ for QCD ones (to be evaluated at some suitable scale, see below). The splitting kernels $P_{a \to bc}(z)$ are \begin{eqnarray} P_{\mathrm{q} \to \mathrm{q}\mathrm{g}}(z)&=&C_F \, \frac{1+z^2}{1-z} ~, \nonumber \\ P_{\mathrm{g} \to \mathrm{g}\g}(z)&=&N_C \, \frac{(1-z(1-z))^2}{z(1-z)} ~, \nonumber \\ P_{\mathrm{g} \to \mathrm{q}\overline{\mathrm{q}}}(z)&=&T_R \, (z^2 + (1-z)^2) ~, \nonumber \\ P_{\mathrm{q} \to \mathrm{q}\gamma}(z)&=&e_{\mathrm{q}}^2 \, \frac{1+z^2}{1-z} ~, \nonumber \\ P_{\ell \to \ell\gamma}(z)&=&e_{\ell}^2 \, \frac{1+z^2}{1-z} ~, \label{sh:APker} \end{eqnarray} with $C_F = 4/3$, $N_C = 3$, $T_R = n_f/2$ (i.e. $T_R$ receives a contribution of $1/2$ for each allowed $\mathrm{q}\overline{\mathrm{q}}$ flavour), and $e_{\mathrm{q}}^2$ and $e_{\ell}^2$ the squared electric charge ($4/9$ for $\u$-type quarks, $1/9$ for $\d$-type ones, and 1 for leptons). Persons familiar with analytical calculations may wonder why the `+ prescriptions' and $\delta(1-z)$ terms of the splitting kernels in eq.~(\ref{sh:APker}) are missing. These complications fulfil the task of ensuring flavour and energy conservation in the analytical equations. The corresponding problem is solved trivially in Monte Carlo programs, where the shower evolution is traced in detail, and flavour and four-momentum are conserved at each branching. The legacy left is the need to introduce a cut-off on the allowed range of $z$ in splittings, so as to avoid the singular regions corresponding to excessive production of very soft gluons. Also note that $P_{\mathrm{g} \to \mathrm{g}\g}(z)$ is given here with a factor $N_C$ in front, while it is sometimes shown with $2 N_C$. The confusion arises because the final state contains two identical partons. With the normalization above, $P_{a \to bc}(z)$ is interpreted as the branching probability for the original parton $a$. On the other hand, one could also write down the probability that a parton $b$ is produced with a fractional energy $z$. Almost all the above kernels can be used unchanged also for this purpose, with the obvious symmetry $P_{a \to bc}(z) = P_{a \to cb}(1-z)$. For $\mathrm{g} \to \mathrm{g}\g$, however, the total probability to find a gluon with energy fraction $z$ is the sum of the probability to find either the first or the second daughter there, and that gives the factor of 2 enhancement. \subsubsection{The Sudakov form factor} \label{ss:sudakov} The $t$ variable fills the function of a kind of time for the shower evolution. In final-state showers, $t$ is constrained to be gradually decreasing away from the hard scattering, in initial-state ones to be gradually increasing towards the hard scattering. This does not mean that an individual parton runs through a range of $t$ values: in the end, each parton is associated with a fixed $t$ value, and the evolution procedure is just a way of picking that value. It is only the ensemble of partons in many events that evolves continuously with $t$, cf. the concept of parton distributions. For a given $t$ value we define the integral of the branching probability over all allowed $z$ values, \begin{equation} {\cal I}_{a \to bc}(t) = \int_{z_{-}(t)}^{z_{+}(t)} \d z \, \frac{\alpha_{abc}}{2 \pi} \, P_{a \to bc}(z) ~. \end{equation} The na\"{\i}ve probability that a branching occurs during a small range of $t$ values, $\delta t$, is given by $\sum_{b,c} {\cal I}_{a \to bc}(t) \, \delta t$, and thus the probability for no emission by $1 - \sum_{b,c} {\cal I}_{a \to bc}(t) \, \delta t$. If the evolution of parton $a$ starts at a `time' $t_0$, the probability that the parton has not yet branched at a `later time' $t > t_0$ is given by the product of the probabilities that it did not branch in any of the small intervals $\delta t$ between $t_0$ and $t$. In other words, letting $\delta t \to 0$, the no-branching probability exponentiates: \begin{equation} {\cal P}_{\mrm{no-branching}}(t_0,t) = \exp \left\{ - \int_{t_0}^t \d t' \, \sum_{b,c} {\cal I}_{a \to bc}(t') \right\} = S_a(t) ~. \label{sh:Pnobranch} \end{equation} Thus the actual probability that a branching of $a$ occurs at $t$ is given by \begin{equation} \frac{\d {\cal P}_a}{\d t} = - \frac{\d {\cal P}_{\mrm{no-branching}}(t_0,t)}{\d t} = \left( \sum_{b,c} {\cal I}_{a \to bc}(t) \right) \exp \left\{ - \int_{t_0}^t \d t' \, \sum_{b,c} {\cal I}_{a \to bc}(t') \right\} ~. \label{sh:Pbranch} \end{equation} The first factor is the na\"{\i}ve branching probability, the second the suppression due to the conservation of total probability: if a parton has already branched at a `time' $t' < t$, it can no longer branch at $t$. This is nothing but the exponential factor that is familiar from radioactive decay. In parton-shower language the exponential factor $S_a(t) = {\cal P}_{\mrm{no-branching}}(t_0,t)$ is referred to as the Sudakov form factor \cite{Sud56}. The ordering in terms of increasing $t$ above is the appropriate one for initial-state showers. In final-state showers the evolution is from an initial $t_{\mathrm{max}}$ (set by the hard scattering) and towards smaller $t$. In that case the integral from $t_0$ to $t$ in eqs.~(\ref{sh:Pnobranch}) and (\ref{sh:Pbranch}) is replaced by an integral from $t$ to $t_{\mathrm{max}}$. Since, by convention, the Sudakov factor is still defined from the lower cut-off $t_0$, i.e. gives the probability that a parton starting at scale $t$ will not have branched by the lower cut-off scale $t_0$, the no-branching factor is actually ${\cal P}_{\mrm{no-branching}}(t_{\mathrm{max}},t) = S_a(t_{\mathrm{max}})/S_a(t)$. We note that the above structure is exactly of the kind discussed in section \ref{ss:vetoalg}. The veto algorithm is therefore extensively used in the Monte Carlo simulation of parton showers. \subsubsection{Matching to the hard scattering} \label{sss:showermatching} The evolution in $Q^2$ is begun from some maximum scale $Q_{\mathrm{max}}^2$ for final-state parton showers, and is terminated at (a possibly different) $Q_{\mathrm{max}}^2$ for initial-state showers. In general $Q_{\mathrm{max}}^2$ is not known. Indeed, since the parton-shower language does not guarantee agreement with higher-order matrix-element results, neither in absolute shape nor normalization, there is no unique prescription for a `best' choice. Generically $Q_{\mathrm{max}}$ should be of the order of the hard-scattering scale, i.e. the largest virtuality should be associated with the hard scattering, and initial- and final-state parton showers should only involve virtualities smaller than that. This may be viewed just as a matter of sound bookkeeping: in a $2 \to n$ graph, a $2 \to 2$ hard-scattering subgraph could be chosen several different ways, but if all the possibilities were to be generated then the cross section would be double-counted. Therefore one should define the $2 \to 2$ `hard' piece of a $2 \to n$ graph as the one that involves the largest virtuality. Of course, the issue of double-counting depends a bit on what processes are actually generated in the program. If one considers a $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ final state in hadron colliders, this could come either as final-state radiation off a $\mathrm{q} \overline{\mathrm{q}}$ pair, or by a gluon splitting in a $\mathrm{q} \overline{\mathrm{q}}$ pair, or many other ways, so that the danger of double-counting is very real. On the other hand, consider the production of a low-$p_{\perp}$, low-mass Drell--Yan pair of leptons, together with two quark jets. Such a process in principle could proceed by having a $\gamma^*$ emitted off a quark leg, with a quark--quark scattering as hard interaction. However, since this process is not included in the program, there is no actual danger of (this particular) double-counting, and so the scale of evolution could be picked larger than the mass of the Drell--Yan pair, at least by some amount. For most $2 \to 2$ scattering processes in {\tsc{Pythia}}, the $Q^2$ scale of the hard scattering is chosen to be $Q_{\mrm{hard}}^2 = p_{\perp}^2$ (when the final-state particles are massless, otherwise masses are added). In final-state showers, where $Q$ is associated with the mass of the branching parton, transverse momenta generated in the shower are constrained by $p_{\perp} < Q/2$. An ordering that shower $p_{\perp}$ be smaller than the hard-scattering $p_{\perp}$ therefore corresponds roughly to $Q_{\mathrm{max}}^2 = 4 Q_{\mrm{hard}}^2$, which is the default assumption. In principle, the constraints are slightly different for initial-state showers, but not enough to warrant a separate $Q_{\mathrm{max}}$ choice. The situation is rather better for the final-state showers in the decay of any colour-singlet particles, such as the $\mathrm{Z}^0$ or the $\H^0$, either as part of a hard $2 \to 1 \to 2$ process, or anywhere else in the final state. Then we know that $Q_{\mathrm{max}}$ has to be put equal to the particle mass. It is also possible to match the parton-shower evolution to the first-order matrix-element results. In the program this is done under the assumption that the resonance has spin one, and this approach is known to work very well for $\gamma^* / \Z^0$. The machinery is not fully correct for the spin-zero $\H^0$, but should also there provide a rather good description. QCD processes such as $\mathrm{q}\mathrm{g} \to \mathrm{q}\mathrm{g}$ pose a special problem when the scattering angle is small. Coherence effects (see below) may then restrict the emission further than what is just given by the $Q_{\mathrm{max}}$ scale introduced above. This is most easily viewed in the rest frame of the $2 \to 2$ hard scattering subprocess. Some colours flow from the initial to the final state. The radiation associated with such a colour flow should be restricted to a cone with opening angle given by the difference between the original and the final colour directions; there is one such cone around the incoming parton for initial state radiation and one around the outgoing parton for final state radiation. Colours that are annihilated or created in the process effectively correspond to an opening angle of 180$^{\circ}$ and therefore the emission is not constrained for these. For a gluon, which have two colours and therefore two different cones, a random choice is made between the two for the first branching. Further, coherence effects also imply azimuthal anisotropies of the emission inside the allowed cones. \subsection{Final-State Showers} Final-state showers are time-like, i.e. all virtualities $m^2 = E^2 - \mbf{p}^2 \geq 0$. The maximum allowed virtuality scale $Q^2_{\mathrm{max}}$ is set by the hard-scattering process, and thereafter the virtuality is decreased in each subsequent branching, down to the cut-off scale $Q_0^2$. This cut-off scale is used to regulate both soft and collinear divergences in the emission probabilities. The main points of the {\tsc{Jetset}} showering algorithm are as follows. \begin{Itemize} \item It is a leading-log algorithm, of the improved, coherent kind, i.e. with angular ordering. \item It can be used for an arbitrary initial pair of partons or, in fact, for any one, two or three given entities (including hadrons and gauge bosons) although only quarks, gluons and leptons can initiate a shower. \item The pair of showering partons may be given in any frame, but the evolution is carried out in the c.m. frame of the showering partons. \item Energy and momentum are conserved exactly at each step of the showering process. \item If the initial pair is $\mathrm{q}\overline{\mathrm{q}}$ or $\ell^+ \ell^-$ (coming from a resonance decay) an additional rejection technique is used in the first branching of each of the two original partons, so as to reproduce the lowest-order differential 3-jet cross section. \item In subsequent branchings, angular ordering (coherence effects) is imposed. \item Gluon helicity effects, i.e. correlations between the production plane and the decay plane of a gluon, can be included. \item The first-order $\alpha_{\mathrm{s}}$ expression is used, with the $Q^2$ scale given by (an approximation to) the squared transverse momentum of a branching. The default $\Lambda_{\mrm{QCD}}$, which should not be regarded as a proper $\Lambda_{\br{\mrm{MS}}}$, is 0.4 GeV. \item The parton shower is by default cut off at a mass scale of 1 GeV. \end{Itemize} Let us now proceed with a more detailed description. \subsubsection{The choice of evolution variable} In the {\tsc{Jetset}} shower algorithm, the evolution variable $Q^2$ is associated with the squared mass of the branching parton, $Q^2 = m_a^2$ for a branching $a \to bc$. As a consequence, $t = \ln(Q^2/\Lambda^2) = \ln(m_a^2/\Lambda^2)$. This $Q^2$ choice is not unique, and indeed other programs have other definitions: \tsc{Herwig} uses $Q^2 \approx m^2/(2z(1-z))$ \cite{Mar88} and \tsc{Ariadne} $Q^2 = p_{\perp}^2 \approx z(1-z)m^2$ \cite{Pet88}. With $Q$ a mass scale, the lower cut-off $Q_0$ is one in mass. To be more precise, in a QCD shower, the $Q_0$ parameter is used to derive effective masses \begin{eqnarray} m_{\mrm{eff},\mathrm{g}} & = & \frac{1}{2} Q_0 ~, \nonumber \\ m_{\mrm{eff},\mathrm{q}} & = & \sqrt{ m_{\mathrm{q}}^2 + \frac{1}{4} Q_0^2 } ~, \end{eqnarray} where the $m_{\mathrm{q}}$ have been chosen as typical current-algebra quark masses. A parton cannot branch unless its mass is at least the sum of the lightest pair of allowed decay products, i.e. the minimum mass scale at which a branching is possible is \begin{eqnarray} m_{\mathrm{min},\mathrm{g}} & = & 2 \, m_{\mrm{eff},\mathrm{g}} = Q_0 ~, \nonumber \\ m_{\mathrm{min},\mathrm{q}} & = & m_{\mrm{eff},\mathrm{q}} + m_{\mrm{eff},\mathrm{g}} \geq Q_0 ~. \end{eqnarray} The above masses are used to constrain the allowed range of $Q^2$ and $z$ values. However, once it has been decided that a parton cannot branch any further, that parton is put on the mass shell, i.e. `final-state' gluons are massless. When also photon emission is included, a separate $Q_0$ scale is introduced for the QED part of the shower, exactly reproducing the QCD one above \cite{Sjo92c}. By default the two $Q_0$ scales are chosen equal, and have the value 1 GeV. If anything, one would be inclined to allow a cut-off lower for photon emission than for gluon one. In that case the allowed $z$ range of photon emission would be larger than that of gluon emission, and at the end of the shower evolution only photon emission would be allowed. Photon and gluon emission differ fundamentally in that photons appear as physical particles in the final state, while gluons are confined. For photon emission off quarks, however, the confinement forces acting on the quark may provide an effective photon emission cut-off at larger scales than the bare quark mass. Soft and collinear photons could also be emitted by the final-state charged hadrons; the matching between emission off quarks and off hadrons is a delicate issue, and we therefore do not attempt to address the soft-photon region. For photon emission off leptons, there is no need to introduce any collinear emission cut-off beyond what is given by the lepton mass, but we keep the same cut-off approach as for quarks: firstly, the program is not aimed at high-precision studies of lepton pairs (where interference terms between initial- and final-state radiation also would have to be included); secondly, most experimental procedures would include the energy of collinear photons into the effective energy of a final-state lepton. \subsubsection{The choice of energy splitting variable} The final-state radiation machinery is always applied in the c.m. frame of the hard scattering, from which normally emerges a pair of evolving partons. Occasionally there may be one evolving parton recoiling against a non-evolving one, as in $\mathrm{q} \overline{\mathrm{q}} \to \mathrm{g} \gamma$, where only the gluon evolves in the final state, but where the energy of the photon is modifed by the branching activity of the gluon. (With only one evolving parton and nothing else, it would not be possible to conserve energy and momentum when the parton is assigned a mass.) Thus, before the evolution is performed, the parton pair is boosted to their common c.m. frame, and rotated to sit along the $z$ axis. After the evolution, the full parton shower is rotated and boosted back to the original frame of the parton pair. The interpretation of the energy and momentum splitting variable $z$ is not unique, and in fact the program allows the possibility to switch between four different alternatives \cite{Ben87a}, `local' and `global' $z$ definition combined with `constrained' or `unconstrained' evolution. In all four of them, the $z$ variable is interpreted as an energy fraction, i.e. $E_b = z E_a$ and $E_c = (1-z) E_a$. In the `local' choice of $z$ definition, energy fractions are defined in the rest frame of the grandmother, i.e. the mother of parton $a$. The preferred choice is the `global' one, in which energies are always evaluated in the c.m. frame of the hard scattering. The two definitions agree for the branchings of the partons that emerge directly from the hard scattering, since the hard scattering itself is considered to be the `mother' of the first generation of partons. For instance, in $\mathrm{Z}^0 \to \mathrm{q}\overline{\mathrm{q}}$ the $\mathrm{Z}^0$ is considered the mother of the $\mathrm{q}$ and $\overline{\mathrm{q}}$, even though the branching is not handled by the parton-showering machinery. The `local' and `global' definitions diverge for subsequent branchings, where the `global' tends to allow more shower evolution. In a branching $a \to bc$ the kinematically allowed range of $z = z_a$ values, $z_{-} < z < z_{+}$, is given by \begin{equation} z_{\pm} = \frac{1}{2} \left\{ 1 + \frac{m_b^2 - m_c^2}{m_a^2} \pm \frac{|\mbf{p}_a|}{E_a} \, \frac{\sqrt{ (m_a^2 - m_b^2 - m_c^2)^2 - 4 m_b^2 m_c^2}}{m_a^2} \right\} ~. \label{sh:zbounds} \end{equation} With `constrained' evolution, these bounds are respected in the evolution. The cut-off masses $m_{\mrm{eff},b}$ and $m_{\mrm{eff},c}$ are used to define the maximum allowed $z$ range, within which $z_a$ is chosen, together with the $m_a$ value. In the subsequent evolution of $b$ and $c$, only pairs of $m_b$ and $m_c$ are allowed for which the already selected $z_a$ fulfils the constraints in eq.~(\ref{sh:zbounds}). For `unconstrained' evolution, which is the preferred alternative, one may start off by assuming the daughters to be massless, so that the allowed $z$ range is \begin{equation} z_{\pm} = \frac{1}{2} \left\{ 1 \pm \frac{|\mbf{p}_a|}{E_a} \theta(m_a - m_{\mathrm{min},a}) \right\} ~, \label{sh:zrange} \end{equation} where $\theta(x)$ is the step function, $\theta(x) = 1$ for $x > 0$ and $\theta(x) = 0$ for $x < 0$. The decay kinematics into two massless four-vectors $p_b^{(0)}$ and $p_c^{(0)}$ is now straightforward. Once $m_b$ and $m_c$ have been found from the subsequent evolution, subject only to the constraints $m_b < z_a E_a$, $m_c < (1-z_a) E_a$ and $m_b + m_c < m_a$, the actual massive four-vectors may be defined as \begin{equation} p_{b,c} = p_{b,c}^{(0)} \pm (r_c p_c^{(0)} - r_b p_b^{(0)}) ~, \label{sh:pshowershift} \end{equation} where \begin{equation} r_{b,c} = \frac{m_a^2 \pm (m_c^2 -m_b^2) - \sqrt{ (m_a^2 - m_b^2 - m_c^2)^2 - 4 m_b^2 m_c^2}}{2 m_a^2} ~. \end{equation} In other words, the meaning of $z_a$ is somewhat reinterpreted {\it post facto}. Needless to say, the `unconstrained' option allows more branchings to take place than the `constrained' one. In the following discussion we will only refer to the `global, unconstrained' $z$ choice. \subsubsection{First branchings and matrix-element matching} The final-state evolution is normally started from some initial parton pair $1 + 2$, at a $Q_{\mathrm{max}}^2$ scale determined by deliberations already discussed. When the evolution of parton 1 is considered, it is assumed that parton 2 is massless, so that the parton 1 energy and momentum are simple functions of its mass (and of the c.m. energy of the pair, which is fixed), and hence also the allowed $z_1$ range for splittings is a function of this mass, eq.~(\ref{sh:zrange}). Correspondingly, parton 2 is evolved under the assumption that parton 1 is massless. After both partons have been assigned masses, their correct energies may be found, which are smaller than originally assumed. Therefore the allowed $z$ ranges have shrunk, and it may happen that a branching has been assigned a $z$ value outside this range. If so, the parton is evolved downwards in mass from the rejected mass value; if both $z$ values are rejected, the parton with largest mass is evolved further. It may also happen that the sum of $m_1$ and $m_2$ is larger than the c.m. energy, in which case the one with the larger mass is evolved downwards. The checking and evolution steps are iterated until an acceptable set of $m_1$, $m_2$, $z_1$ and $z_2$ has been found. The procedure is an extension of the veto algorithm, where an initial overestimation of the allowed $z$ range is compensated by rejection of some branchings. One should note, however, that the veto algorithm is not strictly applicable for the coupled evolution in two variables ($m_1$ and $m_2$), and that therefore some arbitrariness is involved. This is manifest in the choice of which parton will be evolved further if both $z$ values are unacceptable, or if the mass sum is too large. For quark and lepton pairs which come from the decay of a colour-singlet particle, the first branchings are matched to the explicit first-order matrix elements for gauge boson decays. This is also done, e.g. in $\H^0$ decays, which has spin 0 rather than 1, and for which in principle therefore the matrix elements are slightly different. The matching is based on a mapping of the parton-shower variables on to the 3-jet phase space. To produce a 3-jet event, $\gamma^* / \Z^0 \to \mathrm{q}(p_1) \overline{\mathrm{q}}(p_2) \mathrm{g}(p_3)$, in the shower language, one will pass through an intermediate state, where either the $\mathrm{q}$ or the $\overline{\mathrm{q}}$ is off the mass shell. If the former is the case then \begin{eqnarray} m^2 & = & (p_1 + p_3)^2 = E_{\mrm{cm}}^2 (1 - x_2) ~, \nonumber \\ z & = & \frac{E_1}{E_1 + E_3} = \frac{x_1}{x_1 + x_3} = \frac{x_1}{2-x_2} ~, \label{sh:MEfrPS} \end{eqnarray} where $x_i = 2 E_i/E_{\mrm{cm}}$. The $\overline{\mathrm{q}}$ emission case is obtained with $1 \leftrightarrow 2$. The parton-shower splitting expression in terms of $m^2$ and $z$, eq.~(\ref{sh:Patobc}), can therefore be translated into the following differential 3-jet rate: \begin{eqnarray} \frac{1}{\sigma} \, \frac{\d \sigma_{\mrm{PS}}}{\d x_1 \, \d x_2} & = & \frac{\alpha_{\mathrm{s}}}{2 \pi} \, C_F \, \frac{1}{(1-x_1)(1-x_2)} \times \nonumber \\ & \times & \left\{ \frac{1-x_1}{x_3} \left( 1 + \left( \frac{x_1}{2-x_2} \right)^2 \right) + \frac{1-x_2}{x_3} \left( 1 + \left( \frac{x_2}{2-x_1} \right)^2 \right) \right\} ~, \label{sh:PSwt} \end{eqnarray} where the first term inside the curly bracket comes from emission off the quark and the second term from emission off the antiquark. The corresponding expression in matrix-element language is \begin{equation} \frac{1}{\sigma} \, \frac{\d \sigma_{\mrm{ME}}}{\d x_1 \, \d x_2} = \frac{\alpha_{\mathrm{s}}}{2 \pi} \, C_F \, \frac{1}{(1-x_1)(1-x_2)} \left\{ x_1^2 +x_2^2 \right\} ~. \label{sh:MEwt} \end{equation} With the kinematics choice of \tsc{Jetset}, the matrix-element expression is always smaller than the parton-shower one. It is therefore possible to run the shower as usual, but to impose an extra weight factor $\d \sigma_{\mrm{ME}} / \d \sigma_{\mrm{PS}}$, which is just the ratio of the expressions in curly brackets. If a branching is rejected, the evolution is continued from the rejected $Q^2$ value onwards (the veto algorithm). The weighting procedure is applied to the first branching of both the $\mathrm{q}$ and the $\overline{\mathrm{q}}$, in each case with the (nominal) assumption that none of the other partons branch (neither the sister nor the daughters), so that the relations of eq.~(\ref{sh:MEfrPS}) are applicable. If a photon is emitted instead of a gluon, the emission rate in parton showers is given by \begin{eqnarray} \frac{1}{\sigma} \, \frac{\d \sigma_{\mrm{PS}}}{\d x_1 \, \d x_2} & = & \frac{\alpha_{\mathrm{em}}}{2 \pi} \, \frac{1}{(1-x_1)(1-x_2)} \times \nonumber \\ & \times & \left\{ e_{\mathrm{q}}^2 \, \frac{1-x_1}{x_3} \left( 1 + \left( \frac{x_1}{2-x_2} \right)^2 \right) + e_{\overline{\mathrm{q}}}^2 \, \frac{1-x_2}{x_3} \left( 1 + \left( \frac{x_2}{2-x_1} \right)^2 \right) \right\} ~, \end{eqnarray} and in matrix elements by \cite{Gro81} \begin{equation} \frac{1}{\sigma} \, \frac{\d \sigma_{\mrm{ME}}}{\d x_1 \, \d x_2} = \frac{\alpha_{\mathrm{em}}}{2 \pi} \, \frac{1}{(1-x_1)(1-x_2)} \left\{ \left( e_{\mathrm{q}} \, \frac{1-x_1}{x_3} - e_{\overline{\mathrm{q}}} \, \frac{1-x_2}{x_3} \right)^2 \left( x_1^2 +x_2^2 \right) \right\} ~. \end{equation} As in the gluon emission case, a weighting factor $\d \sigma_{\mrm{ME}} / \d \sigma_{\mrm{PS}}$ can therefore be applied when either the original $\mathrm{q}$ ($\ell$) or the original $\overline{\mathrm{q}}$ ($\br{\ell}$) emits a photon. For a neutral resonance, such as $\mathrm{Z}^0$, where $e_{\overline{\mathrm{q}}} = - e_{\mathrm{q}}$, the above expressions simplify and one recovers exactly the same ratio $\d \sigma_{\mrm{ME}} / \d \sigma_{\mrm{PS}}$ as for gluon emission. Compared with the standard matrix-element treatment, a few differences remain. The shower one automatically contains the Sudakov form factor and an $\alpha_{\mathrm{s}}$ running as a function of the $p_{\perp}^2$ scale of the branching. The shower also allows all partons to evolve further, which means that the na\"{\i}ve kinematics assumed for a comparison with matrix elements is modified by subsequent branchings, e.g. that the energy of parton 1 is reduced when parton 2 is assigned a mass. All these effects are formally of higher order, and so do not affect a first-order comparison. This does not mean that the corrections need be small, but experimental results are encouraging: the approach outlined does every bit as good as explicit second-order matrix elements for the description of 4-jet production. \subsubsection{Subsequent branches and angular ordering} The shower evolution is (almost) always done on a pair of partons, so that energy and momentum can be conserved. In the first step of the evolution, the two original partons thus undergo branchings $1 \to 3 + 4$ and $2 \to 5 + 6$. As described above, the allowed $m_1$, $m_2$, $z_1$ and $z_2$ ranges are coupled by kinematical constraints. In the second step, the pair $3 + 4$ is evolved and, separately, the pair $5 + 6$. Considering only the former (the latter is trivially obtained by symmetry), the partons thus have nominal initial energies $E_3^{(0)} = z_1 E_1$ and $E_4^{(0)} = (1-z_1) E_1$, and maximum allowed virtualities $m_{\mathrm{max},3} = \min(m_1,E_3^{(0)})$ and $m_{\mathrm{max},4} = \min(m_1,E_4^{(0)})$. Initially partons 3 and 4 are evolved separately, giving masses $m_3$ and $m_4$ and splitting variables $z_3$ and $z_4$. If $m_3 + m_4 > m_1$, the parton of 3 and 4 that has the largest ratio of $m_i/m_{\mathrm{max},i}$ is evolved further. Thereafter eq.~(\ref{sh:pshowershift}) is used to construct corrected energies $E_3$ and $E_4$, and the $z$ values are checked for consistency. If a branching has to be rejected because the change of parton energy puts $z$ outside the allowed range, the parton is evolved further. This procedure can then be iterated for the evolution of the two daughters of parton 3 and for the two of parton 4, etc., until each parton reaches the cut-off mass $m_{\mathrm{min}}$. Then the parton is put on the mass shell. The model, as described so far, produces so-called conventional showers, wherein masses are strictly decreasing in the shower evolution. Emission angles are decreasing only in an average sense, however, which means that also fairly `late' branchings can give partons at large angles. Theoretical studies beyond the leading-log level show that this is not correct \cite{Mue81}, but that destructive interference effects are large in the region of non-ordered emission angles. To a very good first approximation, these so-called coherence effects can be taken into account in parton shower programs by requiring a strict ordering in terms of decreasing emission angles. The coherence phenomenon is known already from QED. One manifestation is the Chudakov effect \cite{Chu55}, discovered in the study of high-energy cosmic $\gamma$ rays impinging on a nuclear target. If a $\gamma$ is converted into a highly collinear $\e^+\e^-$ pair inside the emulsion, the $\mathrm{e}^+$ and $\mathrm{e}^-$ in their travel through the emulsion ionize atoms and thereby produce blackening. However, near the conversion point the blackening is small: the $\mathrm{e}^+$ and $\mathrm{e}^-$ then are still close together, so that an atom traversed by the pair does not resolve the individual charges of the $\mathrm{e}^+$ and the $\mathrm{e}^-$, but only feels a net charge close to zero. Only later, when the $\mathrm{e}^+$ and $\mathrm{e}^-$ are separated by more than a typical atomic radius, are the two able to ionize independently of each other. The situation is similar in QCD, but is further extended, since now also gluons carry colour. For example, in a branching $\mathrm{q}_0 \to \mathrm{q}\mathrm{g}$ the $\mathrm{q}$ and $\mathrm{g}$ share a colour--anticolour pair, and therefore the $\mathrm{q}$ and $\mathrm{g}$ cannot emit subsequent gluons incoherently. Again the net effect is to reduce the amount of soft gluon emission: since a soft gluon (emitted at large angles) corresponds to a large (transverse) wavelength, the soft gluon is unable to resolve the separate colour charges of the $\mathrm{q}$ and the $\mathrm{g}$, and only feels the net charge carried by the $\mathrm{q}_0$. Such a soft gluon $\mathrm{g}'$ (in the region $\theta_{\mathrm{q}_0 \mathrm{g}'} > \theta_{\mathrm{q} \mathrm{g}}$) could therefore be thought of as being emitted by the $\mathrm{q}_0$ rather than by the $\mathrm{q}$--$\mathrm{g}$ system. If one considers only emission that should be associated with the $\mathrm{q}$ or the $\mathrm{g}$, to a good approximation, there is a complete destructive interference in the regions of non-decreasing opening angles, while partons radiate independently of each other inside the regions of decreasing opening angles ($\theta_{q g'} < \theta_{q g}$ and $\theta_{g g'} < \theta_{q g}$), once azimuthal angles are averaged over. The details of the colour interference pattern are reflected in non-uniform azimuthal emission probabilities. The first branchings of the shower are not affected by the angular-ordering requirement --- since the evolution is performed in the c.m. frame of the original parton pair, where the original opening angle is 180$^{\circ}$, any angle would anyway be smaller than this --- but here instead the matrix-element matching procedure is used, where applicable. Subsequently, each opening angle is compared with that of the preceding branching in the shower. For a branching $a \to bc$ the kinematical approximation \begin{equation} \theta_a \approx \frac{p_{\perp b}}{E_b} + \frac{p_{\perp c}}{E_c} \approx \sqrt{z_a (1-z_a)} m_a \left( \frac{1}{z_a E_a} + \frac{1}{(1-z_a) E_a} \right) = \frac{1}{\sqrt{z_a(1-z_a)}} \frac{m_a}{E_a} \end{equation} is used to derive the opening angle (this is anyway to the same level of approximation as the one in which angular ordering is derived). With $\theta_b$ of the $b$ branching calculated similarly, the requirement $\theta_b < \theta_a$ can be reduced to \begin{equation} \frac{z_b (1-z_b)}{m_b^2} > \frac{1-z_a}{z_a m_a^2} ~. \end{equation} Since photons do not obey angular ordering, the check on angular ordering is not performed when a photon is emitted. When a gluon is emitted in the branching after a photon, its emission angle is restricted by that of the preceding QCD branching in the shower, i.e. the photon emission angle does not enter. \subsubsection{Other final-state shower aspects} The electromagnetic coupling constant for the emission of photons on the mass shell is $\alpha_{\mathrm{em}} = \alpha_{\mathrm{em}}(Q^2 = 0) \approx 1/137$. For the strong coupling constant several alternatives are available, the default being the first-order expression $\alpha_{\mathrm{s}}(p_{\perp}^2)$, where $p_{\perp}^2$ is defined by the approximate expression $p_{\perp}^2 \approx z(1-z) m^2$. Studies of next-to-leading-order corrections favour this choice \cite{Ama80}. The other alternatives are a fixed $\alpha_{\mathrm{s}}$ and an $\alpha_{\mathrm{s}}(m^2)$. With the default choice of $p_{\perp}^2$ as scale in $\alpha_{\mathrm{s}}$, a further cut-off is introduced on the allowed phase space of gluon emission, not present in the options with fixed $\alpha_{\mathrm{s}}$ or with $\alpha_{\mathrm{s}}(m^2)$, nor in the QED shower. A minimum requirement, to ensure a well-defined $\alpha_{\mathrm{s}}$, is that $p_{\perp} / \Lambda > 1.1$, but additionally {\tsc{Jetset}} requires that $p_{\perp} > Q_0/2$. This latter requirement is not a necessity, but it makes sense when $p_{\perp}$ is taken to be the preferred scale of the branching process, rather than e.g. $m$. It reduces the allowed $z$ range, compared with the purely kinematical constraints. Since the $p_{\perp}$ cut is not present for photon emission, the relative ratio of photon to gluon emission off a quark is enhanced at small virtualities compared with na\"{\i}ve expectations; in actual fact this enhancement is largely compensated by the running of $\alpha_{\mathrm{s}}$, which acts in the opposite direction. The main consequence, however, is that the gluon energy spectrum is peaked at around $Q_0$ and rapidly vanishes for energies below that, whilst the photon spectum extends all the way to zero energy. Previously it was said that azimuthal angles in branchings are chosen isotropically. In fact, as an option, it is possible to include some effects of gluon polarization, which correlate the production and the decay planes of a gluon, such that a $\mathrm{g} \to \mathrm{g}\g$ branching tends to take place in the production plane of the gluon, while a decay out of the plane is favoured for $\mathrm{g} \to \mathrm{q}\overline{\mathrm{q}}$. The formulae are given e.g. in ref. \cite{Web86}, as simple functions of the $z$ value at the vertex where the gluon is produced and of the $z$ value when it branches. Also coherence phenomena lead to non-isotropic azimuthal distributions \cite{Web86}, which are included as a further option. In either case the $\varphi$ azimuthal variable is first chosen isotropically, then the weight factor due to polarization times coherence is evaluated, and the $\varphi$ value is accepted or rejected. In case of rejection, a new $\varphi$ is generated, and so on. While the rule is to have an initial pair of partons, there are a few examples where one or three partons have to be allowed to shower. If only one parton is given, it is not possible to conserve both energy and momentum. The choice has been made to conserve energy and jet direction, but the momentum vector is scaled down when the radiating parton acquires a mass. The `rest frame of the system', used e.g. in the $z$ definition, is taken to be whatever frame the jet is given in. In $\Upsilon \to \mathrm{g}\g\mathrm{g}$ decays and other primary three-parton configurations, one is left with the issue how the energy sharing variables $x_1$ and $x_2$ from the massless matrix elements should be reinterpreted for a massive three-parton configuration. We have made the arbitrary choice of preserving the energy of each parton, which means that relative angles between the original partons is changed. Mass triplets outside the allowed phase space are rejected and the evolution continued. Finally, it should be noted that two toy shower models are included as options. One is a scalar gluon model, in which the $\mathrm{q} \to \mathrm{q}\mathrm{g}$ branching kernel is replaced by $P_{\mathrm{q} \to \mathrm{q}\mathrm{g}}(z) = \frac{2}{3} (1-z)$. The couplings of the gluon, $\mathrm{g} \to \mathrm{g}\g$ and $\mathrm{g} \to \mathrm{q}\overline{\mathrm{q}}$, have been left as free parameters, since they depend on the colour structure assumed in the model. The spectra are flat in $z$ for a spin 0 gluon. Higher-order couplings of the type $\mathrm{g} \to \mathrm{g}\g\mathrm{g}$ could well contribute significantly, but are not included. The second toy model is an Abelian vector one. In this option $\mathrm{g} \to \mathrm{g} \mathrm{g}$ branchings are absent, and $\mathrm{g} \to \mathrm{q} \overline{\mathrm{q}}$ ones enhanced. More precisely, in the splitting kernels, eq. (\ref{sh:APker}), the Casimir factors are changed as follows: $C_F = 4/3 \to 1$, $N_C = 3 \to 0$, $T_R = n_f/2 \to 3n_f$. When using either of these options, one should be aware that also a number of other components in principle should be changed, from the running of $\alpha_{\mathrm{s}}$ to the whole concept of fragmentation. One should therefore not take them too seriously. \subsection{Initial-State Showers} The initial-state showe algorithm in {\tsc{Pythia}} is not quite as sophisticated as the final-state one. This is partly because initial-state radiation is less well understood theoretically, partly because the programming task is more complicated and ambiguous. Still, the program at disposal is known to do a reasonably good job of describing existing data, such as $\mathrm{Z}^0$ production properties at hadron colliders \cite{Sjo85}. \subsubsection{The shower structure} \label{sss:initshowstruc} A fast hadron may be viewed as a cloud of quasireal partons. Similarly a fast lepton may be viewed as surrounded by a cloud of photons and partons; in the program the two situations are on an equal footing, but here we choose the hadron as example. At each instant, an individual parton can initiate a virtual cascade, branching into a number of partons. This cascade can be described in terms of a tree-like structure, composed of many subsequent branchings $a \to bc$. Each branching involves some relative transverse momentum between the two daughters. In a language where four-momentum is conserved at each vertex, this implies that at least one of the $b$ and $c$ partons must have a space-like virtuality, $m^2 < 0$. Since the partons are not on the mass shell, the cascade only lives a finite time before reassembling, with those parts of the cascade that are most off the mass shell living the shortest time. A hard scattering, e.g. in deep inelastic leptoproduction, will probe the hadron at a given instant. The probe, i.e. the virtual photon in the leptoproduction case, is able to resolve fluctuations in the hadron up to the $Q^2$ scale of the hard scattering. Thus probes at different $Q^2$ values will seem to see different parton compositions in the hadron. The change in parton composition with $t = \ln(Q^2/\Lambda^2)$ is given by the evolution equations \begin{equation} \frac{\d f_b(x,t)}{\d t} = \sum_{a,c} \int \frac{\d x'}{x'} \, f_a(x',t) \, \frac{\alpha_{abc}}{2 \pi} \, P_{a \to bc} \left( \frac{x}{x'} \right) ~. \label{sh:sfevol} \end{equation} Here the $f_i(x,t)$ are the parton-distribution functions, expressing the probability of finding a parton $i$ carrying a fraction $x$ of the total momentum if the hadron is probed at virtuality $Q^2$. The $P_{a \to bc}(z)$ are given in eq.~(\ref{sh:APker}). As before, $\alpha_{abc}$ is $\alpha_{\mathrm{s}}$ for QCD shower and $\alpha_{\mathrm{em}}$ for QED ones. Eq.~(\ref{sh:sfevol}) is closely related to eq.~(\ref{sh:Patobc}): $\d {\cal P}_a$ describes the probability that a given parton $a$ will branch (into partons $b$ and $c$), $\d f_b$ the influx of partons $b$ from the branchings of partons $a$. (The expression $\d f_b$ in principle also should contain a loss term for partons $b$ that branch; this term is important for parton-distribution evolution, but does not appear explicitly in what we shall be using eq. (\ref{sh:sfevol}) for.) The absolute form of hadron parton distributions cannot be predicted in perturbative QCD, but rather have to be parametrized at some $Q_0$ scale, with the $Q^2$ dependence thereafter given by eq.~(\ref{sh:sfevol}). Available parametrizations are discussed in section \ref{ss:structfun}. The lepton and photon parton distributions inside a lepton can be fully predicted, but here for simplicity are treated on equal footing with hadron parton distributions. If a hard interaction scatters a parton out of the incoming hadron, the `coherence' \cite{Gri83} of the cascade is broken: the partons can no longer reassemble completely back to the cascade-initiating parton. In this semiclassical picture, the partons on the `main chain' of consecutive branchings that lead directly from the initiating parton to the scattered parton can no longer reassemble, whereas fluctuations on the `side branches' to this chain may still disappear. A convenient description is obtained by assigning a space-like virtuality to the partons on the main chain, in such a way that the partons on the side branches may still be on the mass shell. Since the momentum transfer of the hard process can put the scattered parton on the mass shell (or even give it a time-like virtuality, so that it can initiate a final-state shower), one is then guaranteed that no partons have a space-like virtuality in the final state. (In real life, confinement effects obviously imply that partons need not be quite on the mass shell.) If no hard scattering had taken place, the virtuality of the space-like parton line would still force the complete cascade to reassemble. Since the virtuality of the cascade probed is carried by one single parton, it is possible to equate the space-like virtuality of this parton with the $Q^2$ scale of the cascade, to be used e.g. in the evolution equations. Further, coherence effects \cite{Gri83,Bas83} guarantee that the $Q^2$ vaules of the partons along the main chain are strictly ordered, with the largest $Q^2$ values close to the hard scattering. In recent years, further coherence effects have been studied \cite{Cia87}, with particular implications for the structure of parton showers at small $x$. None of these additional complications are implemented in the current algorithm, with the exception of a few rather primitive options that do not address the full complexity of the problem. Instead of having a tree-like structure, where all legs are treated democratically, the cascade is reduced to a single sequence of branchings $a \to bc$, where the $a$ and $b$ partons are on the main chain of space-like virtuality, $m_{a,b}^2 < 0$, while the $c$ partons are on the mass shell and do not branch. (Later we will include the possibility that the $c$ partons may have positive virtualities, $m_c^2 > 0$, which leads to the appearance of time-like `final-state' parton showers on the side branches.) This truncation of the cascade is only possible when it is known which parton actually partakes in the hard scattering: of all the possible cascades that exist virtually in the incoming hadron, the hard scattering will select one. To obtain the correct $Q^2$ evolution of parton distributions, e.g., it is essential that all branches of the cascade be treated democratically. In Monte Carlo simulation of space-like showers this is a major problem. If indeed the evolution of the complete cascade is to be followed from some small $Q_0^2$ up to the $Q^2$ scale of the hard scattering, it is no possible at the same time to handle kinematics exactly, since the virtuality of the various partons cannot be found until after the hard scattering has been selected. This kind of `forward evolution' scheme therefore requires a number of extra tricks to be made to work. Further, in this approach it is not known e.g. what the $\hat{s}$ of the hard scattering subsystem will be until the evolution has been carried out, which means that the initial-state evolution and the hard scattering have to be selected jointly, a not so trivial task. Instead we use the `backwards evolution' approach \cite{Sjo85}, in which the hard scattering is first selected, and the parton shower that preceded it is subsequently reconstructed. This reconstruction is started at the hard interaction, at the $Q_{\mathrm{max}}^2$ scale, and thereafter step by step one moves `backwards' in `time', towards smaller $Q^2$, all the way back to the parton-shower initiator at the cut-off scale $Q_0^2$. This procedure is possible if evolved parton distributions are used to select the hard scattering, since the $f_i(x,Q^2)$ contain the inclusive summation of all initial-state parton-shower histories that can lead to the appearance of an interacting parton $i$ at the hard scale. What remains is thus to select an exclusive history from the set of inclusive ones. \subsubsection{Longitudinal evolution} The evolution equations, eq.~(\ref{sh:sfevol}), express that, during a small increase $\d t$ there is a probability for parton $a$ with momentum fraction $x'$ to become resolved into parton $b$ at $x = z x'$ and another parton $c$ at $x' - x = (1-z) x'$. Correspondingly, in backwards evolution, during a decrease $\d t$ a parton $b$ may be `unresolved' into parton $a$. The relative probability $\d {\cal P}_b$ for this to happen is given by the ratio $\d f_b / f_b$. Using eq.~(\ref{sh:sfevol}) one obtains \begin{equation} \d {\cal P}_b = \frac{\d f_b(x,t)}{f_b(x,t)} = |\d t| \, \sum_{a,c} \int \frac{\d x'}{x'} \, \frac{f_a(x',t)}{f_b(x,t)} \, \frac{\alpha_{abc}}{2 \pi} \, P_{a \to bc} \left( \frac{x}{x'} \right) ~. \end{equation} Summing up the cumulative effect of many small changes $\d t$, the probability for no radiation exponentiates. Therefore one may define a form factor \begin{eqnarray} S_b(x,t_{\mathrm{max}},t) & = & \exp \left\{ - \int_t^{t_{\mathrm{max}}} \d t' \, \sum_{a,c} \int \frac{\d x'}{x'} \, \frac{f_a(x',t')}{f_b(x,t')} \, \frac{\alpha_{abc}(t')}{2\pi} \, P_{a \to bc} \left( \frac{x}{x'} \right) \right\} \nonumber \\ & = & \exp \left\{ - \int_t^{t_{\mathrm{max}}} \d t' \, \sum_{a,c} \int \d z \, \frac{\alpha_{abc}(t')}{2\pi} \, P_{a \to bc}(z) \, \frac{x'f_a(x',t')}{xf_b(x,t')} \right\} ~, \end{eqnarray} giving the probability that a parton $b$ remains at $x$ from $t_{\mathrm{max}}$ to a $t < t_{\mathrm{max}}$. It may be useful to compare this with the corresponding expression for forward evolution, i.e. with $S_a(t)$ in eq.~(\ref{sh:Pnobranch}). The most obvious difference is the appearance of parton distributions in $S_b$. Parton distributions are absent in $S_a$: the probability for a given parton $a$ to branch, once it exists, is independent of the density of partons $a$ or $b$. The parton distributions in $S_b$, on the other hand, express the fact that the probability for a parton $b$ to come from the branching of a parton $a$ is proportional to the number of partons $a$ there are in the hadron, and inversely proportional to the number of partons $b$. Thus the numerator $f_a$ in the exponential of $S_b$ ensures that the parton composition of the hadron is properly reflected. As an example, when a gluon is chosen at the hard scattering and evolved backwards, this gluon is more likely to have been emitted by a $\u$ than by a $\d$ if the incoming hadron is a proton. Similarly, if a heavy flavour is chosen at the hard scattering, the denominator $f_b$ will vanish at the $Q^2$ threshold of the heavy-flavour production, which means that the integrand diverges and $S_b$ itself vanishes, so that no heavy flavour remain below threshold. Another difference between $S_b$ and $S_a$, already touched upon, is that the $P_{\mathrm{g} \to \mathrm{g}\g}(z)$ splitting kernel appears with a normalization $2 N_C$ in $S_b$ but only with $N_C$ in $S_a$, since two gluons are produced but only one decays in a branching. A knowledge of $S_b$ is enough to reconstruct the parton shower backwards. At each branching $a \to bc$, three quantities have to be found: the $t$ value of the branching (which defines the space-like virtuality $Q_b^2$ of parton $b$), the parton flavour $a$ and the splitting variable $z$. This information may be extracted as follows: \begin{Enumerate} \item If parton $b$ partook in the hard scattering or branched into other partons at a scale $t_{\mathrm{max}}$, the probability that $b$ was produced in a branching $a \to bc$ at a lower scale $t$ is \begin{equation} \frac{\d {\cal P}_b}{\d t} = - \frac{\d S_b(x,t_{\mathrm{max}},t)}{\d t} = \left( \sum_{a,c} \int dz \, \frac{\alpha_{abc}(t')}{2\pi} \, P_{a \to bc}(z) \, \frac{x'f_a(x',t')}{xf_b(x,t')} \right) S_b(x,t_{\mathrm{max}},t) ~. \end{equation} If no branching is found above the cut-off scale $t_0$ the iteration is stopped and parton $b$ is assumed to be massless. \item Given the $t$ of a branching, the relative probabilities for the different allowed branchings $a \to bc$ are given by the $z$ integrals above, i.e. by \begin{equation} \int \d z \, \frac{\alpha_{abc}(t)}{2\pi} \, P_{a \to bc}(z) \, \frac{x'f_a(x',t)}{xf_b(x,t)} ~. \label{sh:Pbspace} \end{equation} \item Finally, with $t$ and $a$ known, the probability distribution in the splitting variable $z = x/x' = x_b/x_a$ is given by the integrand in eq.~(\ref{sh:Pbspace}). \end{Enumerate} In addition, the azimuthal angle $\varphi$ of the branching is selected isotropically, i.e. no spin or coherence effects are included in this distribution. The selection of $t$, $a$ and $z$ is then a standard task of the kind than can be performed with the help of the veto algorithm. Specifically, upper and lower bounds for parton distributions are used to find simple functions that are everywhere larger than the integrands in eq.~(\ref{sh:Pbspace}). Based on these simple expressions, the integration over $z$ may be carried out, and $t$, $a$ and $z$ values selected. This set is then accepted with a weight given by a ratio of the correct integrand in eq.~(\ref{sh:Pbspace}) to the simple approximation used, both evaluated for the given set. Since parton distributions, as a rule, are not in a simple analytical form, it may be tricky to find reasonably good bounds to parton distributions. It is necessary to make different assumptions for valence and sea quarks, and be especially attentive close to a flavour threshold (\cite{Sjo85}). An electron distribution inside an electron behaves differently from parton distributions encountered in hadrons, and has to be considered separately. A comment on soft gluon emission. Nominally the range of the $z$ integral in $S_b$ is $x \leq z \leq 1$. The lower limit corresponds to $x' = x/z = 1$, and parton distributions vanish in this limit, wherefore no problems are encountered here. At the upper cut-off $z=1$ the splitting kernels $P_{\mathrm{q} \to \mathrm{q}\mathrm{g}}(z)$ and $P_{\mathrm{g} \to \mathrm{g}\g}$ diverge. This is the soft gluon singularity: the energy carried by the emitted gluon is vanishing, $x_{\mathrm{g}} = x' - x = (1-z) x' = (1-z) x/z \to 0$ for $z \to 1$. In order to calculate the integral over $z$ in $S_b$, an upper cut-off $z_{\mathrm{max}} = x/(x + x_{\epsilon})$ is introduced, i.e. only branchings with $z \leq z_{\mathrm{max}}$ are included in $S_b$. Here $x_{\epsilon}$ is a small number, typically chosen so that the gluon energy $x_{\mathrm{g}} \sqrt{s}/2 \geq x_{\epsilon} \sqrt{s}/2 = 2$ GeV. The average amount of energy carried away by gluons in the range $x_{g} < x_{\epsilon}$, over the given range of $t$ values from $t_a$ to $t_b$, may be estimated \cite{Sjo85}. The finally selected $z$ value may thus be picked as $z = z_{\mrm{hard}} \langle z_{\mrm{soft}}(t_a, t_b) \rangle$, where $z_{\mrm{hard}}$ is the originally selected $z$ value and $z_{\mrm{soft}}$ is the correction factor for soft gluon emission. In QED showers, the smallness of $\alpha_{\mathrm{em}}$ means that one can use rather smaller cut-off values without obtaining large amounts of emission. A fixed small cut-off $x_{\gamma} > 10^{-6}$ is therefore used to avoid the region of very soft photons. As has been discussed in section \ref{sss:estructfun}, the electron distribution inside the electron is cut off at $x_{\mathrm{e}} < 1 - 10^{-6}$, for numerical reasons, so the two cuts are closely matched. The cut-off scale $Q_0$ may be chosen separately for QCD and QED showers, just as in final-state radiation. The defaults are 1 GeV and 0.001 GeV, respectively. The former is the typical hadronic mass scale, below which radiation is not expected resolvable; the latter is of the order of the electron mass. Normally QED and QCD showers do not appear mixed. The most notable exception is resolved photoproduction (in $\e\p$) and resolved 2$\gamma$ events (in $\e^+\e^-$), i.e. shower histories of the type $\mathrm{e} \to \gamma \to \mathrm{q}$. Here the $Q^2$ scales need not be ordered at the interface, i.e. the last $\mathrm{e} \to \mathrm{e}\gamma$ branching may well have a larger $Q^2$ than the first $\mathrm{q} \to \mathrm{q} \mathrm{g}$ one, and the branching $\gamma \to \mathrm{q}$ does not even have a strict parton-shower interpretation for the vector dominance model part of the photon parton distribution. These issues are currently not addressed in full. Rather, based on the $x$ selected for the parton (quark or gluon) at the hard scattering, the $x_{\gamma}$ is selected once and for all in the range $x < x_{\gamma} <1$, according to the distribution implied by eq.~(\ref{pg:foldqgine}). The QCD parton shower is then traced backwards from the hard scattering to the QCD shower initiator at $t_0$. No attempt is made to perform the full QED shower, but rather the beam remnant treatment (see section \ref{ss:beamrem}) is used to find the $\overline{\mathrm{q}}$ (or $\mathrm{g}$) remnant that matches the $\mathrm{q}$ (or $\mathrm{g}$) QCD shower initiator, with the electron itself considered as a second beam remnant. \subsubsection{Transverse evolution} \label{sss:initshowtrans} We have above seen that two parton lines may be defined, stretching back from the hard scattering to the initial incoming hadron wavefunctions at small $Q^2$. Specifically, all parton flavours $i$, virtualities $Q^2$ and energy fractions $x$ may be found. The exact kinematical interpretation of the $x$ variable is not unique, however. For partons with small virtualities and transverse momenta, essentially all definitions agree, but differences may appear for branchings close to the hard scattering. In first-order QED \cite{Ber85} and in some simple QCD toy models \cite{Got86}, one may show that the `correct' choice is the `$\hat{s}$ approach'. Here one requires that $\hat{s} = x_1 x_2 s$, both at the hard scattering scale and at any lower scale, i.e. $\hat{s}(Q^2) = x_1(Q^2) \, x_2(Q^2) \, s$, where $x_1$ and $x_2$ are the $x$ values of the two resolved partons (one from each incoming beam particle) at the given $Q^2$ scale. In practice this means that, at a branching with the splitting variable $z$, the total $\hat{s}$ has to be increased by a factor $1/z$ in the backwards evolution. It also means that branchings on the two incoming legs have to be interleaved in a single monotonic sequence of $Q^2$ values of branchings. For a reconstruction of the complete kinematics in this approach, one should start with the hard scattering, for which $\hat{s}$ has been chosen according to the hard scattering matrix element. By backwards evolution, the virtualities $Q_1^2 = -m_1^2$ and $Q_2^2 = -m_2^2$ of the two interacting partons are reconstructed. Initially the two partons are considered in their common c.m. frame, coming in along the $\pm z$ directions. Then the four-momentum vectors have the non-vanishing components \begin{eqnarray} E_{1,2} & = & \frac{ \hat{s} \pm (Q_2^2 - Q_1^2)}{2 \sqrt{\hat{s}}} ~, \nonumber \\ p_{z1} = - p_{z2} & = & \sqrt{ \frac{ (\hat{s} + Q_1^2 + Q_2^2 )^2 - 4 Q_1^2 Q_2^2 }{ 4 \hat{s} } } ~, \end{eqnarray} with $(p_1 + p_2)^2 = \hat{s}$. If, say, $Q_1^2 > Q_2^2$, then the branching $3 \to 1 + 4$, which produced parton 1, is the one that took place closest to the hard scattering, and the one to be reconstructed first. With the four-momentum $p_3$ known, $p_4 = p_3 - p_1$ is automatically known, so there are four degrees of freedom. One corresponds to a trivial azimuthal angle around the $z$ axis. The $z$ splitting variable for the $3 \to 1 + 4$ vertex is found as the same time as $Q_1^2$, and provides the constraint $(p_3 + p_2)^2 = \hat{s}/z$. The virtuality $Q_3^2$ is given by backwards evolution of parton 3. One degree of freedom remains to be specified, and this is related to the possibility that parton 4 initiates a time-like parton shower, i.e. may have a non-zero mass. The maximum allowed squared mass $m_{\mathrm{max},4}^2$ is found for a collinear branching $3 \to 1 + 4$. In terms of the combinations \begin{eqnarray} s_1 & = & \hat{s} + Q_2^2 + Q_1^2 ~, \nonumber \\ s_3 & = & \frac{\hat{s}}{z} + Q_2^2 + Q_3^2 ~, \nonumber \\ r_1 & = & \sqrt{s_1^2 - 4 Q_2^2 Q_1^2} ~, \nonumber \\ r_3 & = & \sqrt{s_3^2 - 4 Q_2^2 Q_3^2} ~, \end{eqnarray} one obtains \begin{equation} m_{\mathrm{max},4}^2 = \frac{s_1 s_3 - r_1 r_3}{2 Q_2^2} - Q_1^2 - Q_3^2 ~, \end{equation} which, for the special case of $Q_2^2 = 0$, reduces to \begin{equation} m_{\mathrm{max},4}^2 = \left\{ \frac{Q_1^2}{z} - Q_3^2 \right\} \left\{ \frac{\hat{s}}{\hat{s} + Q_1^2} - \frac{\hat{s}}{\hat{s}/z + Q_3^2} \right\} ~. \label{sh:zrangespace} \end{equation} These constraints on $m_4$ are only the kinematical ones, in addition coherence phenomena could constrain the $m_{\mathrm{max},4}$ values further. Some options of this kind are available; the default one is to require additionally that $m_4^2 \leq Q_1^2$, i.e. lesser than the space-like virtuality of the sister parton. With the maximum virtuality given, the final-state showering machinery may be used to give the development of the subsequent cascade, including the actual mass $m_4^2$, with $0 \leq m_4^2 \leq m_{\mathrm{max},4}^2$. The evolution is performed in the c.m. frame of the two `resolved' partons, i.e. that of partons 1 and 2 for the branching $3 \to 1 + 4$, and parton 4 is assumed to have a nominal energy $E_{\mrm{nom},4} = (1/z - 1) \sqrt{\hat{s}}/2$. (Slight modifications appear if parton 4 has a non-vanishing mass $m_{\mathrm{q}}$ or $m_{\ell}$.) Using the relation $m_4^2 = (p_3 - p_1)^2$, the momentum of parton 3 may now be found as \begin{eqnarray} E_3 & = & \frac{1}{2 \sqrt{\hat{s}} } \left\{ \frac{\hat{s}}{z} + Q_2^2 - Q_1^2 - m_4^2 \right\} ~, \nonumber \\ p_{z3} & = & \frac{1}{2 p_{z1}} \left\{ s_3 - 2 E_2 E_3 \right\} ~, \nonumber \\ p_{\perp ,3}^2 & = & \left\{ m_{\mathrm{max},4}^2 - m_4^2 \right\} \, \frac{ (s_1 s_3 + r_1 r_3)/2 - Q_2^2 (Q_1^2 + Q_3^2 + m_4^2)}{r_1^2} ~. \end{eqnarray} The requirement that $m_4^2 \geq 0$ (or $\geq m_f^2$ for heavy flavours) imposes a constraint on allowed $z$ values. This constraint cannot be included in the choice of $Q_1^2$, where it logically belongs, since it also depends on $Q_2^2$ and $Q_3^2$, which are unknown at this point. It is fairly rare (in the order of 10\% of all events) that an unallowed $z$ value is generated, and when it happens it is almost always for one of the two branchings closest to the hard interaction: for $Q_2^2 = 0$ eq.~(\ref{sh:zrangespace}) may be solved to yield $z \leq \hat{s}/(\hat{s} + Q_1^2 - Q_3^2)$, which is a more severe cut for $\hat{s}$ small and $Q_1^2$ large. Therefore an essentially bias-free way of coping is to redo completely any initial-state cascade for which this problem appears. This completes the reconstruction of the $3 \to 1 + 4$ vertex. The subsystem made out of partons 3 and 2 may now be boosted to its rest frame and rotated to bring partons 3 and 2 along the $\pm z$ directions. The partons 1 and 4 now have opposite and compensating transverse momenta with respect to the event axis. When the next vertex is considered, either the one that produces parton 3 or the one that produces parton 2, the 3--2 subsystem will fill the function the 1--2 system did above, e.g. the r\^ole of $\hat{s} = \hat{s}_{12}$ in the formulae above is now played by $\hat{s}_{32} = \hat{s}_{12}/z$. The internal structure of the 3--2 system, i.e. the branching $3 \to 1 + 4$, appears nowhere in the continued description, but has become `unresolved'. It is only reflected in the successive rotations and boosts performed to bring back the new endpoints to their common rest frame. Thereby the hard scattering subsystem 1--2 builds up a net transverse momentum and also an overall rotation of the hard scattering subsystem. After a number of steps, the two outermost partons have virtualities $Q^2 < Q_0^2$ and then the shower is terminated and the endpoints assigned $Q^2 = 0$. Up to small corrections from primordial $k_{\perp}$, discussed in section \ref{ss:beamrem}, a final boost will bring the partons from their c.m. frame to the overall c.m. frame, where the $x$ values of the outermost partons agree also with the light-cone definition. \subsubsection{Other initial-state shower aspects} In the formulae above, $Q^2$ has been used as argument for $\alpha_{\mathrm{s}}$, and not only as the space-like virtuality of partons. This is one possibility, but in fact loop calculations tend to indicate that the proper argument for $\alpha_{\mathrm{s}}$ is not $Q^2$ but $p_{\perp}^2 = (1-z) Q^2$ \cite{Bas83}. The variable $p_{\perp}$ does have the interpretation of transverse momentum, although it is only exactly so for a branching $a \to bc$ with $a$ and $c$ massless and $Q^2 = - m_b^2$, and with $z$ interpreted as light-cone fraction of energy and momentum. The use of $\alpha_{\mathrm{s}}((1-z)Q^2)$ is default in the program. Indeed, if one wanted to, the complete shower might be interpreted as an evolution in $p_{\perp}^2$ rather than in $Q^2$. As we see, the initial-state showering algorithm leads to a net boost and rotation of the hard scattering subsystems. The overall final state is made even more complex by the additional final-state radiation. In principle, the complexity is very physical, but it may still have undesirable side effects. One such, discussed further in section \ref{ss:PYswitchkin}, is that it is very difficult to generate events that fulfill specific kinematics conditions, since kinematics is smeared and even, at times, ambiguous. A special case is encountered in deep inelastic scattering in $\e\p$ collisions. Here the DIS $x$ and $Q^2$ values are defined in terms of the scattered electron direction and energy, and therefore are unambiguous (except for issues of final-state photon radiation close to the electron direction). Neither initial- nor final-state showers preserve the kinematics of the scattered electron, however, and hence the DIS $x$ and $Q^2$ are changed. In principle, this is perfectly legitimate, with the caveat that one then also should use different sets of parton distributions than ones derived from DIS, since these are based on the kinematics of the scattered lepton and nothing else. Alternatively, one might consider showering schemes that leave $x$ and $Q^2$ unchanged. In \cite{Ben88} detailed modifications are presented that make a preservation possible when radiation off the incoming and outgoing electron is neglected, but these are not included in the current version of {\tsc{Pythia}}. What is available, as an option, is a simple machinery which preserves $x$ and $Q^2$ from the effects of QCD radiation, and also from those of primordial $k_{\perp}$ and the beam remnant treatment, as follows. After the showers have been generated, the four-momentum of the scattered lepton is changed to the expected one, based on the nominal $x$ and $Q^2$ values. The azimuthal angle of the lepton is maintained when the transverse momentum is adjusted. Photon radiation off the lepton leg is not fully accounted for, i.e. it is assumed that the energy of final-state photons is added to that of the scattered electron for the definition of $x$ and $Q^2$ (this is the normal procedure for parton-distribution definitions). The change of three-momentum on the lepton side of the event is balanced by the final state partons on the hadron side, excluding the beam remnant but including all the partons both from initial- and final-state showering. The fraction of three-momentum shift taken by each parton is proportional to its original light-cone momentum in the direction of the incoming lepton, i.e. to $E \mp p_z$ for a hadron moving in the $\pm$ direction. This procedure guarantees momentum but not energy conservation. For the latter, one additional degree of freedom is needed, which is taken to be the longitudinal momentum of the initial state shower initiator. As this momentum is modified, the change is shared by the final state partons on the hadron side, according to the same light-cone fractions as before (based on the original momenta). Energy conservation requires that the total change in final state parton energies plus the change in lepton side energy equals the change in initiator energy. This condition can be turned into an iterative procedure to find the initiator momentum shift. Sometimes the procedure may break down. For instance, an initiator with $x > 1$ may be reconstructed. If this should happen, the $x$ and $Q^2$ values of the event are preserved, but new initial and final state showers are generated. After five such failures, the event is completely discared in favour of a new kinematical setup. Kindly note that the four-momentum of intermediate partons in the shower history are not being adjusted. In a listing of the complete event history, energy and momentum need then not be conserved in shower branchings. This mismatch could be fixed up, if need be. The scheme presented above should not be taken too literally, but is rather intended as a contrast to the more sophisticated schemes already on the market, if one would like to understand whether the kind of conservation scheme chosen does affect the observable physics. \subsection{Routines and Common Block Variables} \label{ss:showrout} In this section we collect information on how to use the initial- and final-state showering routines. Of these \ttt{LUSHOW} for final-state radiation is the more generally interesting, since it can be called to let a user-defined parton configuration shower. \ttt{PYSSPA}, on the other hand, is so intertwined with the general structure of a {\tsc{Pythia}} event that it is of little use as a stand-alone product. \drawbox{CALL LUSHOW(IP1,IP2,QMAX)}\label{p:LUSHOW} \begin{entry} \itemc{Purpose:} to generate time-like parton showers, conventional or coherent. The performance of the program is regulated by the switches \ttt{MSTJ(40) - MSTJ(50)} and parameters \ttt{PARJ(81) - PARJ(89)}. In order to keep track of the colour flow information, the positions \ttt{K(I,4)} and \ttt{K(I,5)} have to be organized properly for showering partons. Inside the {\tsc{Jetset/Pythia}} programs, this is done automatically, but for external use proper care must be taken. \iteme{IP1 > 0, IP2 = 0 :} generate a time-like parton shower for the parton in line \ttt{IP1} in common block \ttt{LUJETS}, with maximum allowed mass \ttt{QMAX}. With only one parton at hand, one cannot simultaneously conserve both energy and momentum: we here choose to conserve energy and jet direction, while longitudinal momentum (along the jet axis) is not conserved. \iteme{IP1 > 0, IP2 > 0 :} generate time-like parton showers for the two partons in lines \ttt{IP1} and \ttt{IP2} in the common block \ttt{LUJETS}, with maximum allowed mass for each parton \ttt{QMAX}. For shower evolution, the two partons are boosted to their c.m. frame. Energy and momentum is conserved for the pair of partons, although not for each individually. One of the two partons may be replaced by a nonradiating particle, such as a photon or a diquark; the energy and momentum of this particle will then be modified to conserve the total energy and momentum. \iteme{IP1 > 0, IP2 < 0 :} generate time-like parton showers for the \ttt{-IP2} (at most 3) partons in lines \ttt{IP1}, \ttt{IP1+1}, \ldots \ttt{IPI-IP2-1} in the common block \ttt{LUJETS}, with maximum allowed mass for each parton \ttt{QMAX}. The actions for \ttt{IP2=-1} and \ttt{IP2=-2} correspond to what is described above, but additionally \ttt{IP2=-3} may be used to generate the evolution starting from three given partons (e.g. in $\Upsilon \to \mathrm{g}\g\mathrm{g}$). Then the three partons are boosted to their c.m. frame, energy is conserved for each parton individually and momentum for the system as a whole. \iteme{QMAX :} the maximum allowed mass of a radiating parton, i.e. the starting value for the subsequent evolution. (In addition, the mass of a single parton may not exceed its energy, the mass of a parton in a system may not exceed the invariant mass of the system.) \end{entry} \boxsep \begin{entry} \iteme{SUBROUTINE PYSSPA(IPU1,IPU2) :}\label{p:PYSSPA} to generate the space-like showers of the initial-state radiation. \end{entry} \drawbox{COMMON/LUDAT1/MSTU(200),PARU(200),MSTJ(200),PARJ(200)} \begin{entry} \itemc{Purpose:} to give access to a number of status codes and parameters which regulate the performance of {\tsc{Jetset}}. Most parameters are described in section \ref{ss:JETswitch}; here only those related to \ttt{LUSHOW} are described. \boxsep \iteme{MSTJ(40) :}\label{p:MSTJ40} (D=0) possibility to suppress the branching probability for a branching $\mathrm{q} \to \mathrm{q}\mathrm{g}$ (or $\mathrm{q} \to \mathrm{q}\gamma$) of a quark produced in the decay of an unstable particle with width $\Gamma$, where this width has to be specified by the user in \ttt{PARJ(89)}. The algorithm used is not exact, but still gives some impression of potential effects. This switch ought to have appeared at the end of the current list of shower switches (after \ttt{MSTJ(50)}), but because of lack of space it appears immediately before. \begin{subentry} \iteme{= 0 :} no suppression, i.e. the standard parton-shower machinery. \iteme{= 1 :} suppress radiation by a factor $\chi(\omega) = \Gamma^2 / (\Gamma^2 + \omega^2)$, where $\omega$ is the energy of the gluon (or photon) in the rest frame of the radiating dipole. Essentially this means that hard radiation with $\omega > \Gamma$ is removed. \iteme{= 2 :} suppress radiation by a factor $1 - \chi(\omega) = \omega^2 / (\Gamma^2 + \omega^2)$, where $\omega$ is the energy of the gluon (or photon) in the rest frame of the radiating dipole. Essentially this means that soft radiation with $\omega < \Gamma$ is removed. \end{subentry} \iteme{MSTJ(41) :} (D=2) type of branchings allowed in shower. \begin{subentry} \iteme{= 0 :} no branchings at all, i.e. shower is switched off. \iteme{= 1 :} QCD type branchings of quarks and gluons. \iteme{= 2 :} also emission of photons off quarks and leptons; the photons are assumed on the mass shell. \iteme{= 10 :} as \ttt{=2}, but enhance photon emission by a factor \ttt{PARJ(84)}. This option is unphysical, but for moderate values, \ttt{PARJ(84)}$\leq 10$, it may be used to enhance the prompt photon signal in $\mathrm{q}\overline{\mathrm{q}}$ events. The normalization of the prompt photon rate should then be scaled down by the same factor. The dangers of an improper use are significant, so do not use this option if you do not know what you are doing. \end{subentry} \iteme{MSTJ(42) :} (D=2) branching mode for time-like showers. \begin{subentry} \iteme{= 1 :} conventional branching, i.e. without angular ordering. \iteme{= 2 :} coherent branching, i.e. with angular ordering. \end{subentry} \iteme{MSTJ(43) :} (D=4) choice of $z$ definition in branching. \begin{subentry} \iteme{= 1 :} energy fraction in grandmother's rest frame (`local, constrained'). \iteme{= 2 :} energy fraction in grandmother's rest frame assuming massless daughters, with energy and momentum reshuffled for massive ones (`local, unconstrained'). \iteme{= 3 :} energy fraction in c.m. frame of the showering partons (`global, constrained'). \iteme{= 4 :} energy fraction in c.m. frame of the showering partons assuming massless daughters, with energy and momentum reshuffled for massive ones (`global, unconstrained'). \end{subentry} \iteme{MSTJ(44) :} (D=2) choice of $\alpha_{\mathrm{s}}$ scale for shower. \begin{subentry} \iteme{= 0 :} fixed at \ttt{PARU(111)} value. \iteme{= 1 :} running with $Q^2 = m^2/4$, $m$ mass of decaying parton, $\Lambda$ as stored in \ttt{PARJ(81)} (natural choice for conventional showers). \iteme{= 2 :} running with $Q^2 = z(1-z)m^2$, i.e. roughly $p_{\perp}^2$ of branching, $\Lambda$ as stored in \ttt{PARJ(81)} (natural choice for coherent showers). \end{subentry} \iteme{MSTJ(45) :} (D=5) maximum flavour that can be produced in shower by $\mathrm{g} \to \mathrm{q} \overline{\mathrm{q}}$; also used to determine the maximum number of active flavours in the $\alpha_{\mathrm{s}}$ factor in parton showers (here with a minimum of 3). \iteme{MSTJ(46) :} (D=3) nonhomogeneous azimuthal distributions in a shower branching. \begin{subentry} \iteme{= 0 :} azimuthal angle is chosen uniformly. \iteme{= 1 :} nonhomogeneous azimuthal angle in gluon decays due to a kinematics-dependent effective gluon polarization. Not meaningful for scalar model, i.e. then same as \ttt{=0}. \iteme{= 2 :} nonhomogeneous azimuthal angle in gluon decay due to interference with nearest neighbour (in colour). Not meaningful for Abelian model, i.e. then same as \ttt{=0}. \iteme{= 3 :} nonhomogeneous azimuthal angle in gluon decay due to both polarization (\ttt{=1}) and interference (\ttt{=2}). Not meaningful for Abelian model, i.e. then same as \ttt{=1}. Not meaningful for scalar model, i.e. then same as \ttt{=2}. \end{subentry} \iteme{MSTJ(47) :} (D=3) corrections to the lowest-order $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}$, $\mathrm{q}\overline{\mathrm{q}}\gamma$, $\ell^+\ell^-\gamma$ or $\ell\nu_{\ell}\gamma$ 3-parton matrix element at the first branching of either initial parton in a shower. \begin{subentry} \iteme{= 0 :} no corrections. \iteme{= 1 :} included whenever scattered partons are $\mathrm{q}\overline{\mathrm{q}}$, $\ell^+\ell^-$ or $\ell\nu_{\ell}$. \iteme{= 2 :} always included when shower starts from two partons. \iteme{= 3 :} as \ttt{=1} except that for massive quarks also the massive matrix element expression is used, eq.~(\ref{ee:threejMEmass}), while \ttt{=1} is always based on massless matrix elements. \iteme{= 4 :} as \ttt{=2} except that for massive quarks also the massive matrix element expression is used, while \ttt{=2} is always based on massless matrix elements. \end{subentry} \iteme{MSTJ(48) :} (D=0) possibility to impose maximum angle for the first branching in a shower. \begin{subentry} \iteme{= 0 :} no explicit maximum angle. \iteme{= 1 :} maximum angle given by \ttt{PARJ(85)} for single showering parton, by \ttt{PARJ(85)} and \ttt{PARJ(86)} for pair of showering partons. \end{subentry} \iteme{MSTJ(49) :} (D=0) possibility to change the branching probabilities according to some alternative toy models (note that the $Q^2$ evolution of $\alpha_{\mathrm{s}}$ may well be different in these models, but that only the \ttt{MSTJ(44)} options are at the disposal of the user). \begin{subentry} \iteme{= 0 :} standard QCD branchings. \iteme{= 1 :} branchings according to a scalar gluon theory, i.e. the splitting kernels in the evolution equations are, with a common factor $\alpha_{\mathrm{s}}/(2\pi)$ omitted, $P_{\mathrm{q} \to \mathrm{q}\mathrm{g}} = (2/3) (1-z)$, $P_{\mathrm{g} \to \mathrm{g}\g} =$ \ttt{PARJ(87)}, $P_{\mathrm{g} \to \mathrm{q}\overline{\mathrm{q}}} =$ \ttt{PARJ(88)} (for each separate flavour). The couplings of the gluon have been left as free parameters, since they depend on the colour structure assumed. Note that, since a spin 0 object decays isotropically, the gluon splitting kernels contain no $z$ dependence. \iteme{= 2 :} branchings according to an Abelian vector gluon theory, i.e. the colour factors are changed (compared with QCD) according to $C_F = 4/3 \to 1$, $N_C = 3 \to 0$, $T_R = 1/2 \to 3$. Note that an Abelian model is not expected to contain any coherence effects between gluons, so that one should normally use \ttt{MSTJ(42)=1} and \ttt{MSTJ(46)=} 0 or 1. Also, $\alpha_{\mathrm{s}}$ is expected to increase with increasing $Q^2$ scale, rather than decrease. No such $\alpha_{\mathrm{s}}$ option is available; the one that comes closest is \ttt{MSTJ(44)=0}, i.e. a fix value. \end{subentry} \iteme{MSTJ(50) :} (D=3) possibility to introduce colour coherence effects in the first branching of a final state shower; mainly of relevance for QCD parton--parton scattering processes. \begin{subentry} \iteme{= 0 :} none. \iteme{= 1 :} impose an azimuthal anisotropy. \iteme{= 2 :} restrict the polar ange of a branching to be smaller than the scattering angle of the relevant colour flow. \iteme{= 3 :} both azimuthal anisotropy and restricted polar angles. \itemc{Note:} for subsequent branchings the (polar) angular ordering is automatic (\ttt{MSTP(42)=2}) and \ttt{MSTJ(46)=3}). \end{subentry} \boxsep \iteme{PARJ(81) :}\label{p:PARJ81} (D=0.29 GeV) $\Lambda$ value in running $\alpha_{\mathrm{s}}$ for parton showers (see \ttt{MSTJ(44)}). This is used in all user calls to \ttt{LUSHOW}, in the $\e^+\e^-$ routines of {\tsc{Jetset}}, and in a {\tsc{Pythia}} (or {\tsc{Jetset}}) resonance decay. It is not intended for other timelike showers in {\tsc{Pythia}}, however, for which \ttt{PARP(72)} is used. \iteme{PARJ(82) :} (D=1.0 GeV) invariant mass cut-off $m_{\mathrm{min}}$ of parton showers, below which partons are not assumed to radiate. For $Q^2 = p_{\perp}^2$ (\ttt{MSTJ(44)=2}) \ttt{PARJ(82)}/2 additionally gives the minimum $p_{\perp}$ of a branching. To avoid infinite $\alpha_{\mathrm{s}}$ values, one must have \ttt{PARJ(82)}$ > 2 \times$\ttt{PARJ(81)} for \ttt{MSTJ(44)}$\geq 1$ (this is automatically checked in the program, with $2.2 \times$\ttt{PARJ(81)} as the lowest value attainable). \iteme{PARJ(83) :} (D=1.0 GeV) invariant mass cut-off $m_{\mathrm{min}}$ used for photon emission in parton showers, below which quarks and leptons are not assumed to radiate. The function of \ttt{PARJ(83)} closely parallels that of \ttt{PARJ(82)} for QCD branchings, but there is a priori no requirement that the two be equal. \iteme{PARJ(84) :} (D=1.) used for option \ttt{MSTJ(41)=10} as a multiplicative factor in the promt photon emission rate in final state parton showers. Unphysical but useful technical trick, so beware! \iteme{PARJ(85), PARJ(86) :} (D=10.,10.) maximum opening angles allowed in the first branching of parton showers; see \ttt{MSTJ(48)}. \iteme{PARJ(87) :} (D=0.) coupling of $\mathrm{g} \to \mathrm{g}\g$ in scalar gluon shower, see \ttt{MSTJ(49)=1}. \iteme{PARJ(88) :} (D=0.) coupling of $\mathrm{g} \to \mathrm{q}\overline{\mathrm{q}}$ in scalar gluon shower (per quark species), see \ttt{MSTJ(49)=1}. \iteme{PARJ(89) :} (D=0. GeV) the width of the unstable particle studied for the \ttt{MSTJ(40) > 0} options; to be set by the user (separately for each \ttt{LUSHOW} call, if need be). \end{entry} \drawbox{COMMON/PYPARS/MSTP(200),PARP(200),MSTI(200),PARI(200)} \begin{entry} \itemc{Purpose:} to give access to status code and parameters which regulate the performance of {\tsc{Pythia}}. Most parameters are described in section \ref{ss:PYswitchpar}; here only those related to \ttt{PYSSPA} and \ttt{LUSHOW} are described. \iteme{MSTP(22) :}\label{p:MSTP22} (D=0) special override of normal $Q^2$ definition used for maximum of parton-shower evolution. This option only affects processes 10 and 83 (deep inelastic scattering) and only in lepton--hadron events. \begin{subentry} \iteme{= 0 :} use the scale as given in \ttt{MSTP(32)}. \iteme{= 1 :} use the DIS $Q^2$ scale, i.e. $-\hat{t}$. \iteme{= 2 :} use the DIS $W^2$ scale, i.e. $(-\hat{t})(1-x)/x$. \iteme{= 3 :} use the DIS $Q \times W$ scale, i.e. $(-\hat{t}) \sqrt{(1-x)/x}$. \iteme{= 4 :} use the scale $Q^2 (1-x) \max(1, \ln(1/x))$, as motivated by first order matrix elements \cite{Ing80,Alt78}. \itemc{Note:} in all of these alternatives, a multiplicative factor is introduced by \ttt{PARP(67)} and \ttt{PARP(71)}, as usual. \end{subentry} \iteme{MSTP(61) :}\label{p:MSTP61} (D=1) master switch for initial-state QCD and QED radiation. \begin{subentry} \iteme{= 0 :} off. \iteme{= 1 :} on. \end{subentry} \iteme{MSTP(62) :} (D=3) level of coherence imposed on the space-like parton-shower evolution. \begin{subentry} \iteme{= 1 :} none, i.e. neither $Q^2$ values nor angles need be ordered. \iteme{= 2 :} $Q^2$ values at branches are strictly ordered, increasing towards the hard interaction. \iteme{= 3 :} $Q^2$ values and opening angles of emitted (on-mass-shell or time-like) partons are both strictly ordered, increasing towards the hard interaction. \end{subentry} \iteme{MSTP(63) :} (D=2) structure of associated time-like showers, i.e. showers initiated by emission off the incoming space-like partons. \begin{subentry} \iteme{= 0 :} no associated showers are allowed, i.e. emitted partons are put on the mass shell. \iteme{= 1 :} a shower may evolve, with maximum allowed time-like virtuality set by the phase space only. \iteme{= 2 :} a shower may evolve, with maximum allowed time-like virtuality set by phase space or by \ttt{PARP(71)} times the $Q^2$ value of the space-like parton created in the same vertex, whichever is the stronger constraint. \end{subentry} \iteme{MSTP(64) :} (D=2) choice of $\alpha_{\mathrm{s}}$ and $Q^2$ scale in space-like parton showers. \begin{subentry} \iteme{= 0 :} $\alpha_{\mathrm{s}}$ is taken to be fix at the value \ttt{PARU(111)}. \iteme{= 1 :} first-order running $\alpha_{\mathrm{s}}$ with argument \ttt{PARP(63)}$Q^2$. \iteme{= 2 :} first-order running $\alpha_{\mathrm{s}}$ with argument \ttt{PARP(64)}$k_{\perp}^2 = $\ttt{PARP(64)}$(1-z)Q^2$. \end{subentry} \iteme{MSTP(65) :} (D=1) treatment of soft gluon emission in space-like parton-shower evolution. \begin{subentry} \iteme{= 0 :} soft gluons are entirely neglected. \iteme{= 1 :} soft gluon emission is resummed and included together with the hard radiation as an effective $z$ shift. \end{subentry} \iteme{MSTP(66) :} (D=1) choice of lower cut-off for initial-state QCD radiation in anomalous photoproduction events (see \ttt{MSTP(14)=3}). \begin{subentry} \iteme{= 0 :} the lower $Q^2$ cutoff is the standard one in \ttt{PARP(62)}$^2$. \iteme{= 1 :} the lower cutoff is the larger of \ttt{PARP(62)}$^2$ and \ttt{VINT(283)} or \ttt{VINT(284)}, where the latter is the virtuality scale of the $\gamma \to \mathrm{q}\overline{\mathrm{q}}$ vertex on the appropriate side of the event. \end{subentry} \iteme{MSTP(67) :} (D=2) possibility to introduce colour coherence effects in the first branching of the backwards evolution of an initial state shower; mainly of relevance for QCD parton--parton scattering processes. \begin{subentry} \iteme{= 0 :} none. \iteme{= 2 :} restrict the polar angle of a branching to be smaller than the scattering angle of the relevant colour flow. \itemc{Note 1:} azimuthal anisotropies have not yet been included. \itemc{Note 2:} for subsequent branchings, \ttt{MSTP(62)=3} is used to restrict the (polar) angular range of branchings. \end{subentry} \iteme{MSTP(71) :} (D=1) master switch for final-state QCD and QED radiation. \begin{subentry} \iteme{= 0 :} off. \iteme{= 1 :} on. \end{subentry} \boxsep \iteme{PARP(61) :}\label{p:PARP61} (D=0.25 GeV) $\Lambda$ value used in space-like parton shower (see \ttt{MSTP(64)}). This value may be overwritten, see \ttt{MSTP(3)}. \iteme{PARP(62) :} (D=1. GeV) effective cut-off $Q$ or $k_{\perp}$ value (see \ttt{MSTP(64)}), below which space-like parton showers are not evolved. \iteme{PARP(63) :} (D=0.25) in space-like shower evolution the virtuality $Q^2$ of a parton is multiplied by \ttt{PARP(63)} for use as a scale in $\alpha_{\mathrm{s}}$ and parton distributions when \ttt{MSTP(64)=1}. \iteme{PARP(64) :} (D=1.) in space-like parton-shower evolution the squared transverse momentum evolution scale $k_{\perp}^2$ is multiplied by \ttt{PARP(64)} for use as a scale in $\alpha_{\mathrm{s}}$ and parton distributions when \ttt{MSTP(64)=2}. \iteme{PARP(65) :} (D=2. GeV) effective minimum energy (in c.m. frame) of time-like or on-shell parton emitted in space-like shower; see also \ttt{PARP(66)}. \iteme{PARP(66) :} (D=0.001) effective lower cut-off on $1-z$ in space-like showers, in addition to the cut implied by \ttt{PARP(65)}. \iteme{PARP(67) :} (D=4.) the $Q^2$ scale of the hard scattering (see \ttt{MSTP(32)}) is multiplied by \ttt{PARP(67)} to define the maximum parton virtuality allowed in space-like showers. This does not apply to $s$-channel resonances, where the maximum virtuality is set by $m^2$. \iteme{PARP(68) :} (D=1E-3) lower $Q$ cut-off for QED space-like showers. \iteme{PARP(71) :} (D=4.) the $Q^2$ scale of the hard scattering (see \ttt{MSTP(32)}) is multiplied by \ttt{PARP(71)} to define the maximum parton virtuality allowed in time-like showers. This does not apply to $s$-channel resonances, where the maximum virtuality is set by $m^2$. \iteme{PARP(72) :} (D=0.25 GeV) $\Lambda$ value used in running $\alpha_{\mathrm{s}}$ for timelike parton showers, except for showers in the decay of a resonance. (Resonance decay, e.g. $\gamma^*/Z^0$ decay, is instead set by \ttt{PARJ(81)}.) \end{entry} \clearpage \section{Beam Remnants and Underlying Events} Each incoming beam particle may leave behind a beam remnant, which does not take part in the initial-state radiation or hard scattering process. If nothing else, the remnants need be reconstructed and connected to the rest of the event. In hadron--hadron collisions, the composite nature of the two incoming beam particles implies the additional possibility that several parton pairs undergo separate hard or semi-hard scatterings, `multiple interactions'. This may give a non-negligible contribution to the `underlying event' structure, and thus to the total multiplicity. Finally, in high-luminosity colliders, it is possible to have several collisions between beam particles in one and the same beam crossing, i.e. pile-up events, which further act to build up the general particle production activity that is to be observed by detectors. These three aspects are described in turn, with emphasis on the middle one, that of multiple interactions within a single hadron--hadron collision. The main reference on the multiple interactions model is \cite{Sjo87a}. \subsection{Beam Remnants} \label{ss:beamrem} The initial-state radiation algorithm reconstructs one shower initiator in each beam. (If initial-state radiation is not included, the initiator is nothing but the incoming parton to the hard interaction.) Together the two initiators delineate an interaction subsystem, which contains all the partons that participate in the initial-state showers, in the hard interaction, and in the final-state showers. Left behind are two beam remnants which, to first approximation, just sail through, unaffected by the hard process. (The issue of additional interactions is covered in the next section.) A description of the beam remnant structure contains a few components. First, given the flavour content of a (colour-singlet) beam particle, and the flavour and colour of the initiator parton, it is possible to reconstruct the flavour and colour of the beam remnant. Sometimes the remnant may be represented by just a single parton or diquark, but often the remnant has to be subdivided into two separate objects. In the latter case it is necessary to share the remnant energy and momentum between the two. Due to Fermi motion inside hadron beams, the initiator parton may have a `primordial $k_{\perp}$' transverse momentum motion, which has to be compensated by the beam remnant. If the remnant is subdivided, there may also be a relative transverse momentum. In the end, total energy and momentum has to be conserved. To first approximation, this is ensured within each remnant separately, but some final global adjustments are necessary to compensate for the primordial $k_{\perp}$ and any effective beam remnant mass. Consider first a proton (or, with trivial modifications, any other baryon or antibaryon). \begin{Itemize} \item If the initiator parton is a $\u$ or $\d$ quark, it is assumed to be a valence quark, and therefore leaves behind a diquark beam remnant, i.e. either a $\u\d$ or a $\u\u$ diquark, in a colour antitriplet state. Relative probabilities for different diquark spins are derived within the context of the non-relativistic {\bf SU(6)} model, i.e. flavour {\bf SU(3)} times spin {\bf SU(2)}. Thus a $\u\d$ is $3/4$ $\u\d_0$ and $1/4$ $\u\d_1$, while a $\u\u$ is always $\u\u_1$. \item An initiator gluon leaves behind a colour octet $\u\u\d$ state, which is subdivided into a colour triplet quark and a colour antitriplet diquark. {\bf SU(6)} gives the appropriate subdivision, $1/2$ of the time into $\u + \u\d_0$, $1/6$ into $\u + \u\d_1$ and $1/3$ into $\d + \u\u_1$. \item A sea quark initiator, such as an $\mathrm{s}$, leaves behind a $\u\u\d\overline{\mathrm{s}}$ four-quark state. The PDG flavour coding scheme and the fragmentation routines do not foresee such a state, so therefore it is subdivided into a meson plus a diquark, i.e. $1/2$ into $\u\overline{\mathrm{s}} + \u\d_0$, $1/6$ into $\u\overline{\mathrm{s}} + \u\d_1$ and $1/3$ into $\d\overline{\mathrm{s}} + \u\u_1$. Once the flavours of the meson are determined, the choice of meson multiplet is performed as in the standard fragmentation description. \item Finally, an antiquark initiator, such as an $\overline{\mathrm{s}}$, leaves behind a $\u\u\d\mathrm{s}$ four-quark state, which is subdivided into a baryon plus a quark. Since, to first approximation, the $\mathrm{s}\overline{\mathrm{s}}$ pair comes from the branching $\mathrm{g} \to \mathrm{s}\overline{\mathrm{s}}$ of a colour octet gluon, the subdivision $\u\u\d + \mathrm{s}$ is not allowed, since it would correspond to a colour-singlet $\mathrm{s}\overline{\mathrm{s}}$. Therefore the subdivision is $1/2$ into $\u\d_0\mathrm{s} + \u$, $1/6$ into $\u\d_1\mathrm{s} + \u$ and $1/3$ into $\u\u_1\mathrm{s} + \d$. A baryon is formed among the ones possible for the given flavour content and diquark spin, according to the relative probabilities used in the fragmentation. One could argue for an additional weighting to count the number of baryon states available for a given diquark plus quark combination, but this has not been included. \end{Itemize} One may note that any $\u$ or $\d$ quark taken out of the proton is automatically assumed to be a valence quark. Clearly this is unrealistic, but not quite as bad as it might seem. In particular, one should remember that the beam remnant scenario is applied to the initial-state shower initiators at a scale of $Q_0 \approx 1$ GeV and at an $x$ value usually much larger than the $x$ at the hard scattering. The sea quark contribution therefore normally is negligible. For a meson beam remnant, the rules are in the same spirit, but somewhat easier, since no diquark or baryons need be taken into account. Thus a valence quark (antiquark) initiator leaves behind a valence antiquark (quark), a gluon initiator leaves behind a valence quark plus a valence antiquark, and a sea quark (antiquark) leaves behind a meson (which contains the partner to the sea parton) plus a valence antiquark (quark). A resolved photon is even simpler than a meson, since one does not have to make the distinction between valence and sea flavour. Thus any quark (antiquark) initiator leaves behind the matching antiquark (quark), and a gluon leaves behind a quark + antiquark pair. The relative quark flavour composition in the latter case is assumed proportional to $e_{\mathrm{q}}^2$ among light flavours, i.e. $2/3$ into $\u + \overline{\mathrm{u}}$, $1/6$ into $\d + \overline{\mathrm{d}}$, and $1/6$ into $\mathrm{s} + \overline{\mathrm{s}}$. If one wanted to, one could also have chosen to represent the remnant by a single gluon. If no initial-state radiation is assumed, an electron (or, in general, a lepton or a neutrino) leaves behind no beam remnant. Also when radiation is included, one would expect to recover a single electron with the full beam energy when the shower initiator is reconstructed. This does not have to happen, e.g. if the initial-state shower is cut off at a non-vanishing scale, such that some of the emission at low $Q^2$ values is not simulated. Further, for purely technical reasons, the distribution of an electron inside an electron, $f_{\mathrm{e}}^{\mathrm{e}}(x,Q^2)$, is cut off at $x = 1 - 10^{-6}$. This means that always, when initial-state radiation is included, a fraction of at least $10^{-6}$ of the beam energy has to be put into one single photon along the beam direction, to represent this not simulated radiation. The physics is here slightly different from the standard beam remnant concept, but it is handled with the same machinery. Beam remnants can also apper when the electron is resolved with the use of parton distributions, but initial-state radiation is switched off. Conceptually, this is a contradiction, since it is the initial-state radiation that builds up the parton distributions, but sometimes the combination is still useful. Finally, since QED radiation has not yet been included in events with resolved photons inside electrons, also in this case effective beam remnants have to be assigned by the program. The beam remnant assignments inside an electron, in either of the cases above, is as follows. \begin{Itemize} \item An $\mathrm{e}^-$ initiator leaves behind a $\gamma$ remnant. \item A $\gamma$ initiator leaves behind an $\mathrm{e}^-$ remnant. \item An $\mathrm{e}^+$ initiator leaves behind an $\mathrm{e}^- + \mathrm{e}^-$ remnant. \item A $\mathrm{q}$ ($\overline{\mathrm{q}}$) initiator leaves behind a $\overline{\mathrm{q}} + \mathrm{e}^-$ ($\mathrm{q} + \mathrm{e}^-$) remnant. \item A $\mathrm{g}$ initiator leaves behind a $\mathrm{g} + \mathrm{e}^-$ remnant. One could argue that, in agreement with the treatment of photon beams above, the remnant should be $\mathrm{q} + \overline{\mathrm{q}} + \mathrm{e}^-$. The program currently does not allow for three beam remnant objects, however. \end{Itemize} By the hard scattering and initial-state radiation machinery, the shower initiator has been assigned some fraction $x$ of the four-momentum of the beam particle, leaving behind $1-x$ to the remnant. If the remnant consists of two objects, this energy and momentum has to be shared, somehow. For an electron, the sharing is given from first principles: if, e.g., the initiator is a $\mathrm{q}$, then that $\mathrm{q}$ was produced in the sequence of branchings $\mathrm{e} \to \gamma \to \mathrm{q}$, where $x_{\gamma}$ is distributed according to the convolution in eq.~(\ref{pg:foldqgine}). Therefore the $\overline{\mathrm{q}}$ remnant takes a fraction $\chi = (x_{\gamma} -x)/(1-x)$ of the total remnant energy, and the $\mathrm{e}$ takes $1 - \chi$. For the other beam remnants, the relative energy-sharing variable $\chi$ is not known from first principles, but picked according to some suitable parametrization. Normally several different options are available, that can be set separately for baryon and meson beams, and for hadron + quark and quark + diquark (or antiquark) remnants. In one extreme are shapes in agreement with na\"{\i}ve counting rules, i.e. where energy is shared evenly between `valence' partons. For instance, ${\cal P}(\chi) = 2 \, (1-\chi)$ for the energy fraction taken by the $\mathrm{q}$ in a $\mathrm{q} + \mathrm{q}\q$ remnant. In the other extreme, an uneven distribution could be used, like in parton distributions, where the quark only takes a small fraction and most is retained by the diquark. The default for a $\mathrm{q} + \mathrm{q}\q$ remnant is of this type, \begin{equation} {\cal P}(\chi) \propto \frac{(1 - \chi)^3} {\sqrt{\chi^2 + c_{\mathrm{min}}^2}} ~, \end{equation} with $c_{\mathrm{min}} = 2 \langle m_{\mathrm{q}} \rangle / E_{\mrm{cm}} = (0.6$ GeV)$/ E_{\mrm{cm}}$ providing a lower cut-off. In general, the more uneven the sharing of the energy, the less the total multiplicity in the beam remnant fragmentation. If no multiple interactions are allowed, a rather even sharing is needed to come close to the experimental multiplicity (and yet one does not quite make it). With an uneven sharing there is room to generate more of the total multiplicity by multiple interactions \cite{Sjo87a}. In a photon beam, with a remnant $\mathrm{q} + \overline{\mathrm{q}}$, the $\chi$ variable is chosen the same way it would have been in a corresponding meson remnant. Before the $\chi$ variable is used to assign remnant momenta, it is also necessary to consider the issue of primordial $k_{\perp}$. The initiator partons are thus assigned each a $k_{\perp}$ value, vanishing for an electron or photon inside an electron, distributed either according to a Gaussian or an exponential shape for a hadron, and according to either of these shapes or a power-like shape for a quark or gluon inside a photon (which may in its turn be inside an electron). The interaction subsystem is boosted and rotated to bring it from the frame assumed so far, with each initiator along the $\pm z$ axis, to one where the initiators have the required primordial $k_{\perp}$ values. The $p_{\perp}$ recoil is taken by the remnant. If the remnant is composite, the recoil is all taken by one of the two, namely the one that, in some imagined perturbative splitting language, is the sister of the initiator parton. For instance, when a gluon is taken out of a proton, the recoil is all taken by the lone quark (i.e. nothing by the diquark), since one could have imagined an earlier branching $\mathrm{q}_0 \to \mathrm{q} \mathrm{g}$, below the shower cut-off scale $Q_0$, with $p_{\perp q_0} = 0$. In addition, however, two beam remnants may be given a relative $p_{\perp}$, which is then always chosen as for $\mathrm{q}_i \overline{\mathrm{q}}_i$ pairs in the fragmentation description. The $\chi$ variable is interpreted as a sharing of light-cone energy and momentum, i.e. $E + p_z$ for the beam moving in the $+z$ direction and $E - p_z$ for the other one. When the two transverse masses $m_{\perp 1}$ and $m_{\perp 2}$ of a composite remnant have been constructed, the total transverse mass can therefore be found as \begin{equation} m_{\perp}^2 = \frac{m_{\perp 1}^2}{\chi} + \frac{m_{\perp 2}^2}{1 - \chi} ~, \end{equation} if remnant 1 is the one that takes the fraction $\chi$. The choice of a light-cone interpretation to $\chi$ means the definition is invariant under longitudinal boosts, and therefore does not depend on the beam energy itself. A $\chi$ value close to the na\"{\i}ve borders 0 or 1 can lead to an unreasonably large remnant $m_{\perp}$. Therefore an additional check is introduced, that the remnant $m_{\perp}$ be smaller than the na\"{\i}ve c.m. frame remnant energy, $(1-x) E_{\mrm{cm}}/2$. If this is not the case, a new $\chi$ and a new relative transverse momentum is selected. Whether there is one remnant parton or two, the transverse mass of the remnant is not likely to agree with $1-x$ times the mass of the beam particle, i.e. it is not going to be possible to preserve the energy and momentum in each remnant separately. One therefore allows a shuffling of energy and momentum between the beam remnants from each of the two incoming beams. This may be achieved by performing a (small) longitudinal boost of each remnant system. Since there are two boost degrees of freedom, one for each remnant, and two constraints, one for energy and one for longitudinal momentum, a solution may be found. Under some circumstances, one beam remnant may be absent or of very low energy, while the other one is more complicated. One example is deep inelastic scattering in $\e\p$ collisions, where the electron leaves no remnant, or maybe only a low-energy photon. It is clearly then not possible to balance the two beam remnants against each other. Therefore, if one beam remnant has an energy below 0.2 of the beam energy, i.e. if the initiator parton has $x > 0.8$, then the two boosts needed to ensure energy and momentum conservation are instead performed on the other remnant and on the interaction subsystem. If there is a low-energy remnant at all then, before that, energy and momentum are assigned to the remnant constituent(s) so that the appropriate light-cone combination $E \pm p_z$ is conserved, but not energy or momentum separately. If both beam remnants have low energy, but both still exist, then the one with lower $m_{\perp} / E$ is the one that will not be boosted. \subsection{Multiple Interactions} \label{ss:multint} In this section we present the model used in {\tsc{Pythia}} to describe the possibility that several parton pairs undergo hard interactions in a hadron--hadron collision, and thereby contribute to the overall event activity, in particular at low $p_{\perp}$. The same model is also used to describe the VMD $\gamma \mathrm{p}$ events, where the photon interacts like a hadron. It should from the onset be made clear that this is not an easy topic. In fact, in the full event generation process, probably no other area is as poorly understood as this one. The whole concept of multiple interactions is very controversial, with contraditory experimental conclusions \cite{AFS87}. The multiple interactions scenario presented here \cite{Sjo87a} was the first detailed model for this kind of physics, and is still one of the very few available. We will present two related but separate scenarios, one `simple' model and one somewhat more sophisticated. In fact, neither of them are all that simple, which may make the models look unattractive. However, the world of hadron physics {\it is} complicated, and if we err, it is most likely in being too unsophisticated. The experience gained with the model(s), in failures as well as successes, could be used as a guideline in the evolution of yet more detailed models. Our basic philosophy will be as follows. The total rate of parton--parton interactions, as a function of the transverse momentum scale $p_{\perp}$, is assumed to be given by perturbative QCD. This is certainly true for reasonably large $p_{\perp}$ values, but here we shall also extend the perturbative parton--parton scattering framework into the low-$p_{\perp}$ region. A regularization of the divergence in the cross section for $p_{\perp} \to 0$ has to be introduced, however, which will provide us with the main free parameter of the model. Since each incoming hadron is a composite object, consisting of many partons, there should exist the possibility of several parton pairs interacting when two hadrons collide. It is not unreasonable to assume that the different pairwise interactions take place essentially independently of each other, and that therefore the number of interactions in an event is given by a Poissonian distribution. This is the strategy of the `simple' scenario. Furthermore, hadrons are not only composite but also extended objects, meaning that collisions range from very central to rather peripheral ones. Reasonably, the average number of interactions should be larger in the former than in the latter case. Whereas the assumption of a Poissonian distribution should hold for each impact parameter separately, the distribution in number of interactions should be widened by the spread of impact parameters. The amount of widening depends on the assumed matter distribution inside the colliding hadrons. In the `complex' scenario, different matter distributions are therefore introduced. \subsubsection{The basic cross sections} The QCD cross section for hard $2 \to 2$ processes, as a function of the $p_{\perp}^2$ scale, is given by \begin{equation} \frac{\d \sigma}{\d p_{\perp}^2} = \sum_{i,j,k} \int \d x_1 \int \d x_2 \int \d \hat{t} \, f_i(x_1,Q^2) \, f_j(x_2,Q^2) \, \frac{\d \hat{\sigma}_{ij}^k}{\d \hat{t}} \, \delta \! \left( p_{\perp}^2 - \frac{\hat{t}\hat{u}}{\hat{s}} \right) ~, \label{mi:sigmapt} \end{equation} cf. section \ref{ss:kinemtwo}. Implicitly, from now on we are assuming that the `hardness' of processes is given by the $p_{\perp}$ scale of the scattering. For an application of the formula above to small $p_{\perp}$ values, a number of caveats could be made. At low $p_{\perp}$, the integrals receive major contributions from the small-$x$ region, where parton distributions are poorly understood theoretically (Regge limit behaviour, dense packing problems etc. \cite{Lev90}) and not yet measured. Different sets of parton distributions can therefore give numerically rather different results for the phenomenology of interest. One may also worry about higher-order corrections to the jet rates ($K$ factors), beyond what is given by parton-shower corrections --- one simple option we allow here is to evaluate $\alpha_{\mathrm{s}}$ of the hard scattering process at an optimized scale, $\alpha_{\mathrm{s}}(0.075p_{\perp}^2)$ \cite{Ell86}. The hard scattering cross section above some given $p_{\perp\mathrm{min}}$ is given by \begin{equation} \sigma_{\mrm{hard}} (p_{\perp\mathrm{min}}) = \int_{p_{\perp\mathrm{min}}^2}^{s/4} \frac{\d \sigma}{\d p_{\perp}^2} \, \d p_{\perp}^2 ~. \label{mi:sigmahard} \end{equation} Since the differential cross section diverges roughly like $\d p_{\perp}^2 / p_{\perp}^4$, $\sigma_{\mrm{hard}}$ is also divergent for $p_{\perp\mathrm{min}} \to 0$. We may compare this with the total inelastic, non-diffractive cross section $\sigma_{\mrm{nd}}(s)$ --- elastic and diffractive events are not the topic of this section. At current collider energies $\sigma_{\mrm{hard}} (p_{\perp\mathrm{min}})$ becomes comparable with $\sigma_{\mrm{nd}}$ for $p_{\perp\mathrm{min}} \approx$1.5--2 GeV. This need not lead to contradictions: $\sigma_{\mrm{hard}}$ does not give the hadron--hadron cross section but the parton--parton one. Each of the incoming hadrons may be viewed as a beam of partons, with the possibility of having several parton--parton interactions when the hadrons pass through each other. In this language, $\sigma_{\mrm{hard}} (p_{\perp\mathrm{min}}) / \sigma_{\mrm{nd}}(s)$ is simply the average number of parton--parton scatterings above $p_{\perp\mathrm{min}}$ in an event, and this number may well be larger than unity. While the introduction of several interactions per event is the natural consequence of allowing small $p_{\perp\mathrm{min}}$ values and hence large $\sigma_{\mrm{hard}}$ ones, it is not the solution of $\sigma_{\mrm{hard}} (p_{\perp\mathrm{min}})$ being divergent for $p_{\perp\mathrm{min}} \to 0$: the average $\hat{s}$ of a scattering decreases slower with $p_{\perp\mathrm{min}}$ than the number of interactions increases, so na\"{\i}vely the total amount of scattered partonic energy becomes infinite. One cut-off is therefore obtained via the need to introduce proper multi-parton correlated parton distributions inside a hadron. This is not a part of the standard perturbative QCD formalism and is therefore not built into eq.~(\ref{mi:sigmahard}). In practice, even correlated parton-distribution functions seems to provide too weak a cut, i.e. one is lead to a picture with too little of the incoming energy remaining in the small-angle beam jet region. A more credible reason for an effective cut-off is that the incoming hadrons are colour neutral objects. Therefore, when the $p_{\perp}$ of an exchanged gluon is made small and the transverse wavelength correspondingly large, the gluon can no longer resolve the individual colour charges, and the effective coupling is decreased. This mechanism is not in contradiction to perturbative QCD calculations, which are always performed assuming scattering of free partons (rather than partons inside hadrons), but neither does present knowledge of QCD provide an understanding of how such a decoupling mechanism would work in detail. In the simple model one makes use of a sharp cut-off at some scale $p_{\perp\mathrm{min}}$, while a more smooth dampening is assumed for the complex scenario. \subsubsection{The simple model} In an event with several interactions, it is convenient to impose an ordering. The logical choice is to arrange the scatterings in falling sequence of $x_{\perp} = 2 p_{\perp} / E_{\mrm{cm}}$. The `first' scattering is thus the hardest one, with the `subsequent' (`second', `third', etc.) successively softer. It is important to remember that this terminology is in no way related to any picture in physical time; we do not know anything about the latter. In principle, all the scatterings that occur in an event must be correlated somehow, na\"{\i}vely by momentum and flavour conservation for the partons from each incoming hadron, less na\"{\i}vely by various quantum mechanical effects. When averaging over all configurations of soft partons, however, one should effectively obtain the standard QCD phenomenology for a hard scattering, e.g. in terms of parton distributions. Correlation effects, known or estimated, can be introduced in the choice of subsequent scatterings, given that the `preceding' (harder) ones are already known. With a total cross section of hard interactions $\sigma_{\mrm{hard}} (p_{\perp\mathrm{min}})$ to be distributed among $\sigma_{\mrm{nd}}(s)$ (non-diffractive, inelastic) events, the average number of interactions per event is just the ratio $\br{n} = \sigma_{\mrm{hard}} (p_{\perp\mathrm{min}}) / \sigma_{\mrm{nd}}(s)$. As a starting point we will assume that all hadron collisions are equivalent (no impact parameter dependence), and that the different parton--parton interactions take place completely independently of each other. The number of scatterings per event is then distributed according to a Poissonian with mean $\br{n}$. A fit to collider multiplicity data gives $p_{\perp\mathrm{min}} \approx 1.6$ GeV, which corresponds to $\br{n} \approx 1$. For Monte Carlo generation of these interactions it is useful to define \begin{equation} f(x_{\perp}) = \frac{1}{\sigma_{\mrm{nd}}(s)} \, \frac{\d \sigma}{\d x_{\perp}} ~, \end{equation} with $\d \sigma / \d x_{\perp}$ obtained by analogy with eq. (\ref{mi:sigmapt}). Then $f(x_{\perp})$ is simply the probability to have a parton--parton interaction at $x_{\perp}$, given that the two hadrons undergo a non-diffractive, inelastic collision. The probability that the hardest interaction, i.e. the one with highest $x_{\perp}$, is at $x_{\perp 1}$, is now given by \begin{equation} f(x_{\perp 1}) \exp \left\{ - \int_{x_{\perp 1}}^1 f(x'_{\perp}) \, \d x'_{\perp} \right\} ~, \label{mi:hardest} \end{equation} i.e. the na\"{\i}ve probability to have a scattering at $x_{\perp 1}$ multiplied by the probability that there was no scattering with $x_{\perp}$ larger than $x_{\perp 1}$. This is the familiar exponential dampening in radioactive decays, encountered e.g. in parton showers in section \ref{ss:sudakov}. Using the same technique as in the proof of the veto algorithm, section \ref{ss:vetoalg}, the probability to have an $i$:th scattering at an $x_{\perp i} < x_{\perp i-1} < \cdots < x_{\perp 1} < 1$ is found to be \begin{equation} f(x_{\perp i}) \, \frac{1}{(i-1)!} \left( \int_{x_{\perp i}}^1 f(x'_{\perp}) \, \d x'_{\perp} \right)^{i-1} \exp \left\{ - \int_{x_{\perp i}}^1 f(x'_{\perp}) \, \d x'_{\perp} \right\} ~. \end{equation} The total probability to have a scattering at a given $x_{\perp}$, irrespectively of it being the first, the second or whatever, obviously adds up to give back $f(x_{\perp})$. The multiple interaction formalism thus retains the correct perturbative QCD expression for the scattering probability at any given $x_{\perp}$. With the help of the integral \begin{equation} F(x_{\perp}) = \int_{x_{\perp}}^1 f(x'_{\perp}) \, \d x'_{\perp} = \frac{1}{\sigma_{\mrm{nd}}(s)} \, \int_{s x_{\perp}^2/4}^{s/4} \frac{\d \sigma}{\d p_{\perp}^2} \, \d p_{\perp}^2 \end{equation} (where we assume $F(x_{\perp}) \to \infty$ for $x_{\perp} \to 0$) and its inverse $F^{-1}$, the iterative procedure to generate a chain of scatterings $1 > x_{\perp 1} > x_{\perp 2} > \cdots > x_{\perp i}$ is given by \begin{equation} x_{\perp i} = F^{-1}(F(x_{\perp i-1}) - \ln R_i) ~. \label{mi:iter} \end{equation} Here the $R_i$ are random numbers evenly distributed between 0 and 1. The iterative chain is started with a fictitious $x_{\perp 0} = 1 $ and is terminated when $x_{\perp i}$ is smaller than $x_{\perp \mathrm{min}} = 2 p_{\perp\mathrm{min}} / E_{\mrm{cm}}$. Since $F$ and $F^{-1}$ are not known analytically, the standard veto algorithm is used to generate a much denser set of $x_{\perp}$ values, whereof only some are retained in the end. In addition to the $p_{\perp}^2$ of an interaction, it is also necessary to generate the other flavour and kinematics variables according to the relevant matrix elements. Whereas the ordinary parton distributions should be used for the hardest scattering, in order to reproduce standard QCD phenomenology, the parton distributions to be used for subsequent scatterings must depend on all preceding $x$ values and flavours chosen. We do not know enough about the hadron wave function to write down such joint probability distributions. To take into account the energy `already' used in harder scatterings, a conservative approach is to evaluate the parton distributions, not at $x_i$ for the $i$:th scattered parton from hadron, but at the rescaled value \begin{equation} x'_i = \frac{x_i}{\sum_{j=1}^{i-1} x_j} ~. \end{equation} This is our standard procedure in the program; we have tried a few alternatives without finding any significantly different behaviour in the final physics. In a fraction $\exp(-F(x_{\perp \mathrm{min}}))$ of the events studied, there will be no hard scattering above $x_{\perp \mathrm{min}}$ when the iterative procedure in eq.~(\ref{mi:iter}) is applied. It is therefore also necessary to have a model for what happens in events with no (semi)hard interactions. The simplest possible way to produce an event is to have an exchange of a very soft gluon between the two colliding hadrons. Without (initially) affecting the momentum distribution of partons, the `hadrons' become colour octet objects rather than colour singlet ones. If only valence quarks are considered, the colour octet state of a baryon can be decomposed into a colour triplet quark and an antitriplet diquark. In a baryon-baryon collision, one would then obtain a two-string picture, with each string stretched from the quark of one baryon to the diquark of the other. A baryon-antibaryon collision would give one string between a quark and an antiquark and another one between a diquark and an antidiquark. In a hard interaction, the number of possible string drawings are many more, and the overall situation can become quite complex when several hard scatterings are present in an event. Specifically, the string drawing now depends on the relative colour arrangement, in each hadron individually, of the partons that are about to scatter. This is a subject about which nothing is known. To make matters worse, the standard string fragmentation description would have to be extended, to handle events where two or more valence quarks have been kicked out of an incoming hadron by separate interactions. In particular, the position of the baryon number would be unclear. We therefore here assume that, following the hardest interaction, all subsequent interactions belong to one of three classes. \begin{Itemize} \item Scatterings of the $\mathrm{g}\g \to \mathrm{g}\g $ type, with the two gluons in a colour-singlet state, such that a double string is stretched directly between the two outgoing gluons, decoupled from the rest of the system. \item Scatterings $\mathrm{g}\g \to \mathrm{g}\g$, but colour correlations assumed to be such that each of the gluons is connected to one of the strings `already' present. Among the different possibilities of connecting the colours of the gluons, the one which minimizes the total increase in string length is chosen. This is in contrast to the previous alternative, which roughly corresponds to a maximization of the extra string length. \item Scatterings $\mathrm{g}\g \to \mathrm{q}\overline{\mathrm{q}}$, with the final pair again in a colour-singlet state, such that a single string is stretched between the outgoing $\mathrm{q}$ and $\overline{\mathrm{q}}$. \end{Itemize} By default, the three possibilities are assumed equally probable. Note that the total jet rate is maintained at its nominal value, i.e. scatterings such as $\mathrm{q}\mathrm{g} \to \mathrm{q}\mathrm{g}$ are included in the cross section, but are replaced by a mixture of $\mathrm{g}\g$ and $\mathrm{q}\overline{\mathrm{q}}$ events for string drawing issues. Only the hardest interaction is guaranteed to give strings coupled to the beam remnants. One should not take this approach to colour flow too seriously --- clearly it is a simplification --- but the overall picture does not tend to be very dependent on the particular choice you make. Since a $\mathrm{g}\g \to \mathrm{g}\g$ or $\mathrm{q}\overline{\mathrm{q}}$ scattering need not remain that if initial- and final-state showers were to be included, radiation is only included for the hardest interaction. In practice, there is no problem: except for the hardest interaction, which can be hard because of experimental trigger conditions, it is unlikely for a parton scattering to be so hard that radiation plays a significant r\^ole. In events with multiple interactions, the beam remnant treatment is slightly modified. First the hard scattering is generated, with its associated initial- and final-state radiation, and next any additional multiple interactions. Only thereafter are beam remnants attached to the initator partons of the hardest scattering, using the same machinery as before, except that the energy and momentum already taken away from the beam remnants also include that of the subsequent interactions. \subsubsection{A model with varying impact parameters} Up to this point, it has been assumed that the initial state is the same for all hadron collisions, whereas in fact each collision also is characterized by a varying impact parameter $b$. Within the classical framework of this paper, $b$ is to be thought of as a distance of closest approach, not as the Fourier transform of the momentum transfer. A small $b$ value corresponds to a large overlap between the two colliding hadrons, and hence an enhanced probability for multiple interactions. A large $b$, on the other hand, corresponds to a grazing collision, with a large probability that no parton--parton interactions at all take place. In order to quantify the concept of hadronic matter overlap, one may assume a spherically symmetric distribution of matter inside the hadron, $\rho(\mbf{x}) \, \d^3 x = \rho(r) \, \d^3 x$. For simplicity, the same spatial distribution is taken to apply for all parton species and momenta. Several different matter distributions have been tried, and are available. We will here concentrate on the most extreme one, a double Gaussian \begin{equation} \rho(r) \propto \frac{1 - \beta}{a_1^3} \, \exp \left\{ - \frac{r^2}{a_1^2} \right\} + \frac{\beta}{a_2^3} \, \exp \left\{ - \frac{r^2}{a_2^2} \right\} ~. \label{mi:doubleGauss} \end{equation} This corresponds to a distribution with a small core region, of radius $a_2$ and containing a fraction $\beta$ of the total hadronic matter, embedded in a larger hadron of radius $a_1$. While it is mathematically convenient to have the origin of the two Gaussians coinciding, the physics could well correspond to having three disjoint core regions, reflecting the presence of three valence quarks, together carrying the fraction $\beta$ of the proton momentum. One could alternatively imagine a hard hadronic core surrounded by a pion cloud. Such details would affect e.g. the predictions for the $t$ distribution in elastic scattering, but are not of any consequence for the current topics. To be specific, the values $\beta = 0.5$ and $a_2/a_1 = 0.2$ have been picked as default values. It should be noted that the overall distance scale $a_1$ never enters in the subsequent calculations, since the inelastic, non-diffractive cross section $\sigma_{\mrm{nd}}(s)$ is taken from literature rather than calculated from the $\rho(r)$. Compared to other shapes, like a simple Gaussian, the double Gaussian tends to give larger fluctuations, e.g. in the multiplicity distribution of minimum bias events: a collision in which the two cores overlap tends to have a strongly increased activity, while ones where they do not are rather less active. One also has a biasing effect: hard processes are more likely when the cores overlap, thus hard scatterings are associated with an enhanced multiple interaction rate. This provides one possible explanation for the experimental `pedestal effect'. For a collision with impact parameter $b$, the time-integrated overlap ${\cal O}(b)$ between the matter distributions of the colliding hadrons is given by \begin{eqnarray} & & {\cal O}(b) \propto \int \d t \int \d^3 x \, \rho(x,y,z) \, \rho(x+b,y,z+t) \nonumber \\ & & \propto \frac{(1 - \beta)^2}{2a_1^2} \exp \left\{ - \frac{b^2}{2a_1^2} \right\} + \frac{2 \beta (1-\beta)}{a_1^2+a_2^2} \exp \left\{ - \frac{b^2}{a_1^2+ a_2^2} \right\} + \frac{\beta^2}{2a_2^2} \exp \left\{ - \frac{b^2}{2a_2^2} \right\} ~. \end{eqnarray} The necessity to use boosted $\rho(\mbf{x})$ distributions has been circumvented by a suitable scale transformation of the $z$ and $t$ coordinates. The overlap ${\cal O}(b)$ is obviously strongly related to the eikonal $\Omega(b)$ of optical models. We have kept a separate notation, since the physics context of the two is slightly different: $\Omega(b)$ is based on the quantum mechanical scattering of waves in a potential, and is normally used to describe the elastic scattering of a hadron-as-a-whole, while ${\cal O}(b)$ comes from a purely classical picture of point-like partons distributed inside the two colliding hadrons. Furthermore, the normalization and energy dependence is differently realized in the two formalisms. The larger the overlap ${\cal O}(b)$ is, the more likely it is to have interactions between partons in the two colliding hadrons. In fact, there should be a linear relationship \begin{equation} \langle \tilde{n}(b) \rangle = k {\cal O}(b) ~, \end{equation} where $\tilde{n} = 0, 1, 2, \ldots$ counts the number of interactions when two hadrons pass each other with an impact parameter $b$. The constant of proportionality, $k$, is related to the parton--parton cross section and hence increases with c.m. energy. For each given impact parameter, the number of interactions is assumed to be distributed according to a Poissonian. If the matter distribution has a tail to infinity (as the double Gaussian does), events may be obtained with arbitrarily large $b$ values. In order to obtain finite total cross sections, it is necessary to assume that each event contains at least one semi-hard interaction. The probability that two hadrons, passing each other with an impact parameter $b$, will actually undergo a collision is then given by \begin{equation} {\cal P}_{\mrm{int}}(b) = 1 - \exp ( - \langle \tilde{n}(b) \rangle ) = 1 - \exp (- k {\cal O}(b) ) ~, \end{equation} according to Poissonian statistics. The average number of interactions per event at impact parameter $b$ is now \begin{equation} \langle n(b) \rangle = \frac{ \langle \tilde{n}(b) \rangle }{{\cal P}_{\mrm{int}}(b)} = \frac{ k {\cal O}(b) }{ 1 - \exp (- k {\cal O}(b) ) } ~, \label{mi:nofb} \end{equation} where the denominator comes from the removal of hadron pairs which pass without colliding, i.e. with $\tilde{n} = 0$. The relationship $\langle n \rangle = \sigma_{\mrm{hard}}/\sigma_{\mrm{nd}}$ was earlier introduced for the average number of interactions per non-diffractive, inelastic event. When averaged over all impact parameters, this relation must still hold true: the introduction of variable impact parameters may give more interactions in some events and less in others, but it does not affect either $\sigma_{\mrm{hard}}$ or $\sigma_{\mrm{nd}}$. For the former this is because the perturbative QCD calculations only depend on the total parton flux, for the latter by construction. Integrating eq.~(\ref{mi:nofb}) over $b$, one then obtains \begin{equation} \langle n \rangle = \frac{ \int \langle n(b) \rangle \, {\cal P}_{\mrm{int}}(b) \, \d^2 b} { \int {\cal P}_{\mrm{int}}(b) \, \d^2 b} = \frac{ \int k {\cal O}(b) \, \d^2 b} { \int \left( 1 - \exp (- k {\cal O}(b) ) \right) \, \d^2 b} = \frac{\sigma_{\mrm{hard}}}{\sigma_{\mrm{nd}}} ~. \label{mi:kfromsigma} \end{equation} For ${\cal O}(b)$, $\sigma_{\mrm{hard}}$ and $\sigma_{\mrm{nd}}$ given, with $\sigma_{\mrm{hard}} / \sigma_{\mrm{nd}} > 1$, $k$ can thus always be found (numerically) by solving the last equality. The absolute normalization of ${\cal O}(b)$ is not interesting in itself, but only the relative variation with impact parameter. It is therefore useful to introduce an `enhancement factor' $e(b)$, which gauges how the interaction probability for a passage with impact parameter $b$ compares with the average, i.e. \begin{equation} \langle \tilde{n}(b) \rangle = k{\cal O}(b) = e(b) \, \langle k{\cal O}(b) \rangle ~. \label{mi:ebenh} \end{equation} The definition of the average $\langle k{\cal O}(b) \rangle$ is a bit delicate, since the average number of interactions per event is pushed up by the requirement that each event contain at least one interaction. However, an exact meaning can be given \cite{Sjo87a}. With the knowledge of $e(b)$, the $f(x_{\perp})$ function of the simple model generalizes to \begin{equation} f(x_{\perp},b) = e(b) \, f(x_{\perp}) ~. \end{equation} The na\"{\i}ve generation procedure is thus to pick a $b$ according to the phase space $\d^2 b$, find the relevant $e(b)$ and plug in the resulting $f(x_{\perp},b)$ in the formalism of the simple model. If at least one hard interaction is generated, the event is retained, else a new $b$ is to be found. This algorithm would work fine for hadronic matter distributions which vanish outside some radius, so that the $\d^2 b$ phase space which needs to be probed is finite. Since this is not true for the distributions under study, it is necessary to do better. By analogy with eq.~(\ref{mi:hardest}), it is possible to ask what the probability is to find the hardest scattering of an event at $x_{\perp 1}$. For each impact parameter separately, the probability to have an interaction at $x_{\perp 1}$ is given by $f(x_{\perp},b)$, and this should be multiplied by the probability that the event contains no interactions at a scale $x'_{\perp} > x_{\perp 1}$, to yield the total probability distribution \begin{eqnarray} \frac{\d {\cal P}_{\mrm{hardest}}}{\d^2 b \, \d x_{\perp 1}} & = & f(x_{\perp 1},b) \, \exp \left\{ - \int_{x_{\perp 1}}^1 f(x'_{\perp},b) \, \d x'_{\perp} \right\} \nonumber \\ & = & e(b) \, f(x_{\perp 1}) \, \exp \left\{ - e(b) \int_{x_{\perp 1}}^1 f(x'_{\perp}) \, \d x'_{\perp} \right\} ~. \end{eqnarray} If the treatment of the exponential is deferred for a moment, the distribution in $b$ and $x_{\perp 1}$ appears in factorized form, so that the two can be chosen independently of each other. In particular, a high-$p_{\perp}$ QCD scattering or any other hard scattering can be selected with whatever kinematics desired for that process, and thereafter assigned some suitable `hardness' $x_{\perp 1}$. With the $b$ chosen according to $e(b) \, \d^2 b$, the neglected exponential can now be evaluated, and the event retained with a probability proportional to it. From the $x_{\perp 1}$ scale of the selected interaction, a sequence of softer $x_{\perp i}$ values may again be generated as in the simple model, using the known $f(x_{\perp},b)$. This sequence may be empty, i.e. the event need not contain any further interactions. It is interesting to understand how the algorithm above works. By selecting $b$ according to $e(b) \, \d^2 b$, i.e. ${\cal O}(b) \, \d^2 b$, the primary $b$ distribution is maximally biased towards small impact parameters. If the first interaction is hard, by choice or by chance, the integral of the cross section above $x_{\perp 1}$ is small, and the exponential close to unity. The rejection procedure is therefore very efficient for all standard hard processes in the program --- one may even safely drop the weighting with the exponential completely. The large $e(b)$ value is also likely to lead to the generation of many further, softer interactions. If, on the other hand, the first interaction is not hard, the exponential is no longer close to unity, and many events are rejected. This pulls down the efficiency for `minimum bias' event generation. Since the exponent is proportional to $e(b)$, a large $e(b)$ leads to an enhanced probability for rejection, whereas the chance of acceptance is larger with a small $e(b)$. Among events where the hardest interaction is soft, the $b$ distribution is therefore biased towards larger values (smaller $e(b)$), and there is a small probability for yet softer interactions. To evaluate the exponential factor, the program pretabulates the integral of $f(x_{\perp})$ at the initialization stage, and further increases the Monte Carlo statistics of this tabulation as the run proceeds. The $x_{\perp}$ grid is concentrated towards small $x_{\perp}$, where the integral is large. For a selected $x_{\perp 1}$ value, the $f(x_{\perp})$ integral is obtained by interpolation. After multiplication by the known $e(b)$ factor, the exponential factor may be found. In this section, nothing has yet been assumed about the form of the $\d \sigma / \d p_{\perp}$ spectrum. Like in the impact parameter independent case, it is possible to use a sharp cut-off at some given $p_{\perp\mathrm{min}}$ value. However, now each event is required to have at least one interaction, whereas before events without interactions were retained and put at $p_{\perp} = 0$. It is therefore aesthetically more appealing to assume a gradual turn-off, so that a (semi)hard interaction can be rather soft part of the time. The matrix elements roughly diverge like $\alpha_{\mathrm{s}}(p_{\perp}^2) \, \d p_{\perp}^2 / p_{\perp}^4$ for $p_{\perp} \to 0$. They could therefore be regularized as follows. Firstly, to remove the $1/p_{\perp}^4$ behaviour, multiply by a factor $p_{\perp}^4 / (p_{\perp}^2 + p_{\perp 0}^2)^2$. Secondly, replace the $p_{\perp}^2$ argument in $\alpha_{\mathrm{s}}$ by $p_{\perp}^2 + p_{\perp 0}^2$ or, with the inclusion of the $K$ factor procedure introduced earlier, replace $0.075 \, p_{\perp}^2$ by $0.075 \, (p_{\perp}^2 + p_{\perp 0}^2)$. With these substitutions, a continuous $p_{\perp}$ spectrum is obtained, stretching from $p_{\perp} = 0$ to $E_{\mrm{cm}}/2$. For $p_{\perp} \gg p_{\perp 0}$ the standard perturbative QCD cross section is recovered, while values $p_{\perp} \ll p_{\perp 0}$ are strongly damped. The $p_{\perp 0}$ scale, which now is the main free parameter of the model, in practice comes out to be of the same order of magnitude as the sharp cut-off $p_{\perp\mathrm{min}}$ did, i.e. 1.5--2 GeV. If gluons with large transverse wavelength decouple because of the colour-singlet nature of hadrons, and if the transverse structure of hadrons is assumed to be energy-independent, it is natural to assume that also $p_{\perp\mathrm{min}}$ and $p_{\perp 0}$ are independent of the c.m. energy of the hadron collision. For the impact parameter independent picture this works out fine, with all events being reduced to low-$p_{\perp}$ two-string ones when the c.m. energy is reduced. In the variable impact parameter picture, the whole formalism only makes sense if $\sigma_{\mrm{hard}} > \sigma_{\mrm{nd}}$, see e.g. eq.~(\ref{mi:kfromsigma}). Since $\sigma_{\mrm{nd}}$ does not vanish with decreasing energy, but $\sigma_{\mrm{hard}}$ would do that for a fixed $p_{\perp 0}$, this means that $p_{\perp 0}$ has to be reduced when the energy is decreased below some given threshold. The more `sophisticated' model of this section therefore makes sense at collider energies, whereas it is not well suited for applications at lower energies. \subsection{Pile-up Events} \label{ss:pileup} In high luminosity colliders, there is a non-negligible probability that one single bunch crossing may produce several separate events, so-called pile-up events. This in particular applies to future $\mathrm{p}\p$ colliders like LHC, but one could also consider e.g. $\e^+\e^-$ colliders with high rates of $\gamma\gamma$ collisions. The program therefore contains an option, currently only applicable to hadron--hadron collisions, wherein several events may be generated and put one after the other in the event record, to simulate the full amount of particle production a detector might be facing. The program needs to know the assumed luminosity per bunch--bunch crossing, expressed in mb$^{-1}$. Multiplied by the cross section for pile-up processes studied, $\sigma_{\mrm{pile}}$, this gives the average number of collisions per beam crossing, $\br{n}$. These pile-up events are taken to be of the minimum bias type, with diffractive and elastic events included or not (and a further subdivision of diffractive events into single and double). This means that $\sigma_{\mrm{pile}}$ may be either $\sigma_{\mrm{tot}}$, $\sigma_{\mrm{tot}} - \sigma_{\mrm{el}}$ or $\sigma_{\mrm{tot}} - \sigma_{\mrm{el}} - \sigma_{\mrm{diffr}}$. Which option to choose depends on the detector: most detectors would not be able to observe elastic $\mathrm{p}\p$ scattering, and therefore it would be superfluous to generate that kind of events. In addition, we allow for the possibility that one interaction may be of a rare kind, selected freely by the user. There is no option to generate two `rare' events in the same crossing; normally the likelihood for that kind of occurences should be small. If only minimum bias type events are generated, i.e. if only one cross section is involved in the problem, then the number of events in a crossing is distributed according to a Poissonian with the average number $\br{n}$ as calculated above. The program actually will simulate only those beam crossings where at least one event occurs, i.e. not consider the fraction $\exp(-\br{n})$ of zero-event crossings. Therefore the actually generated average number of pile-up events is $\langle n \rangle = \br{n}/(1-\exp(-\br{n}))$. Now instead consider the other extreme, where one event is supposed be rare, with a cross section $\sigma_{\mrm{rare}}$ much smaller than $\sigma_{\mrm{pile}}$, i.e. $f \equiv \sigma_{\mrm{rare}}/\sigma_{\mrm{pile}} \ll 1$. The probability that a bunch crossing will give $i$ events, whereof one of the rare kind, now is \begin{equation} {\cal P}_i = f \, i \, \exp(-\br{n}) \, \frac{\br{n}^i}{i!} = f \, \br{n} \exp(-\br{n}) \, \frac{\br{n}^{i-1}}{(i-1)!} ~. \end{equation} The na\"{\i}ve Poissonian is suppressed by a factor $f$, since one of the events is rare rather than of the normal kind, but enhanced by a factor $i$, since any one of the $i$ events may be the rare one. As the equality shows, the probability distribution is now a Poissonian in $i-1$: in a beam crossing which produces one rare event, the multiplicity of additional pile-up events is distributed according to a Poissonian with average number $\br{n}$. The total average number of events thus is $\langle n \rangle = \br{n} + 1$. Clearly, for processes with intermediate cross sections, $\br{n} \, \sigma_{\mrm{rare}}/\sigma_{\mrm{pile}} \simeq 1$, also the average number of events will be intermediate, and it is not allowed to assume only one event to be of the `rare' type. We do not consider that kind of situations. Kindly note that, in the current implementation, all events are supposed to be produced at the same vertex (the origin). To simulate the spatial extent of the colliding beams, you would have to assign interaction points yourself, and then shift each event separately by the required amount in space and time. When the pile-up option is used, one main limitation is that event records may become very large when several events are put one after the other, so that the space limit in the \ttt{LUJETS} common block is reached. It is possible to expand the dimension of the common block, see \ttt{MSTU(4)} and \ttt{MSTU(5)}, but only up to about 20\,000 entries, which might not always be enough. For practical reasons, the program will only allow a $\br{n}$ up to 120. The multiplicity distribution is truncated above 200, or when the probability for a multiplicity has fallen below $10^{-6}$, whichever occurs sooner. Also low multiplicities with probabilities below $10^{-6}$ are truncated. \subsection{Common Block Variables} \label{ss:multintpar} Of the routines used to generate beam remnants, multiple interactions and pile-up events, none are intended to be used in standalone mode. The only way to regulate these aspects is therefore via the variables in the \ttt{PYPARS} common block. \drawbox{COMMON/PYPARS/MSTP(200),PARP(200),MSTI(200),PARI(200)} \begin{entry} \itemc{Purpose:} to give access to a number of status codes and parameters which regulate the performance of {\tsc{Pythia}}. Most parameters are described in section \ref{ss:PYswitchpar}; here only those related to beam remnants, multiple interactions and pile-up events are described. If the default values, below denoted by (D=\ldots), are not satisfactory, they must in general be changed before the \ttt{PYINIT} call. Exceptions, i.e. variables which can be changed for each new event, are denoted by (C). \iteme{MSTP(81) :}\label{p:MSTP81} (D=1) master switch for multiple interactions. \begin{subentry} \iteme{= 0 :} off. \iteme{= 1 :} on. \end{subentry} \iteme{MSTP(82) :} (D=1) structure of multiple interactions. For QCD processes, used down to $p_{\perp}$ values below $p_{\perp\mathrm{min}}$, it also affects the choice of structure for the one hard/semi-hard interaction. \begin{subentry} \iteme{= 0 :} simple two-string model without any hard interactions. \iteme{= 1 :} multiple interactions assuming the same probability in all events, with an abrupt $p_{\perp\mathrm{min}}$ cut-off at \ttt{PARP(81)}. \iteme{= 2 :} multiple interactions assuming the same probability in all events, with a continuous turn-off of the cross section at $p_{\perp 0} = $\ttt{PARP(82)}. \iteme{= 3 :} multiple interactions assuming a varying impact parameter and a hadronic matter overlap consistent with a Gaussian matter distribution, with a continuous turn-off of the cross section at $p_{\perp 0} = $\ttt{PARP(82)}. \iteme{= 4 :} multiple interactions assuming a varying impact parameter and a hadronic matter overlap consistent with a double Gaussian matter distribution given by \ttt{PARP(83)} and \ttt{PARP(84)}, with a continuous turn-off of the cross section at $p_{\perp 0} = $\ttt{PARP(82)}. \itemc{Note 1:} For \ttt{MSTP(82)}$\geq 2$ and \ttt{CKIN(3)}$ > $\ttt{PARP(82)}, cross sections given with \ttt{PYSTAT(1)} may be somewhat too large, since (for reasons of efficiency) the probability factor that the hard interaction is indeed the hardest in the event is not included in the cross sections. It is included in the event selection, however, so the events generated are correctly distributed. For \ttt{CKIN(3)} values a couple of times larger than \ttt{PARP(82)} this ceases to be a problem. \itemc{Note 2:} The \ttt{PARP(81)} and, in particular, \ttt{PARP(82)} values are sensitive to the choice of parton distributions, $\Lambda_{\mrm{QCD}}$, etc., in the sense that a change in the latter variables leads to a net change in the multiple interaction rate, which has to be compensated by a retuning of \ttt{PARP(81)} or \ttt{PARP(82)} if one wants to keep the net multiple interaction structure the same. The default \ttt{PARP(81)} value is consistent with the other default values give, i.e. CTEQ 2L parton distributions etc. When options \ttt{MSTP(82)=} 2--4 are used, the default \ttt{PARP(82)} value is to be used in conjunction with \ttt{MSTP(2)=2} and \ttt{MSTP(33)=3}. These switches should be set by you. \end{subentry} \iteme{MSTP(83) :} (D=100) number of Monte Carlo generated phase-space points per bin (whereof there are 20) in the initialization (in \ttt{PYMULT}) of multiple interactions for \ttt{MSTP(82)}$\geq 2$. \iteme{MSTP(91) :} (D=1) (C) primordial $k_{\perp}$ distribution in hadron. See \ttt{MSTP(93)} for photon. \begin{subentry} \iteme{= 0 :} no primordial $k_{\perp}$. \iteme{= 1 :} Gaussian, width given in \ttt{PARP(91)}, upper cut-off in \ttt{PARP(93)}. \iteme{= 2 :} exponential, width given in \ttt{PARP(92)}, upper cut-off in \ttt{PARP(93)}. \end{subentry} \iteme{MSTP(92) :} (D=4) (C) energy partitioning in hadron or resolved photon remnant, when this remnant is split into two jets. (For a splitting into a hadron plus a jet, see \ttt{MSTP(94)}.) The energy fraction $\chi$ taken by one of the two objects, with conventions as described for \ttt{PARP(94)} and \ttt{PARP(96)}, is chosen according to the different distributions below. Here $c_{\mathrm{min}} = 2 \langle m_{\mathrm{q}} \rangle/E_{\mrm{cm}} = 0.6$ GeV$/E_{\mrm{cm}}$. \begin{subentry} \iteme{= 1 :} 1 for meson or resolved photon, $2(1-\chi)$ for baryon, i.e. simple counting rules. \iteme{= 2 :} $(k+1)(1-\chi)^k$, with $k$ given by \ttt{PARP(94)} or \ttt{PARP(96)}. \iteme{= 3 :} proportional to $(1-\chi)^k/\sqrt[4]{\chi^2+c_{\mathrm{min}}^2}$, with $k$ given by \ttt{PARP(94)} or \ttt{PARP(96)}. \iteme{= 4 :} proportional to $(1-\chi)^k/\sqrt{\chi^2+c_{\mathrm{min}}^2}$, with $k$ given by \ttt{PARP(94)} or \ttt{PARP(96)}. \iteme{= 5 :} proportional to $(1-\chi)^k/(\chi^2+c_{\mathrm{min}}^2)^{b/2}$, with $k$ given by \ttt{PARP(94)} or \ttt{PARP(96)}, and $b$ by \ttt{PARP(98)}. \end{subentry} \iteme{MSTP(93) :} (D=1) (C) primordial $k_{\perp}$ distribution in photon, either it is one of the incoming particles or inside an electron. \begin{subentry} \iteme{= 0 :} no primordial $k_{\perp}$. \iteme{= 1 :} Gaussian, width given in \ttt{PARP(99)}, upper cut-off in \ttt{PARP(100)}. \iteme{= 2 :} exponential, width given in \ttt{PARP(99)}, upper cut-off in \ttt{PARP(100)}. \iteme{= 3 :} power-like of the type $\d k_{\perp}^2/(k_{\perp 0}^2 + k_{\perp}^2)^2$, with $k_{\perp 0}$ in \ttt{PARP(99)} and upper $k_{\perp}$ cut-off in \ttt{PARP(100)}. \iteme{= 4 :} power-like of the type $\d k_{\perp}^2/(k_{\perp 0}^2 + k_{\perp}^2)$, with $k_{\perp 0}$ in \ttt{PARP(99)} and upper $k_{\perp}$ cut-off in \ttt{PARP(100)}. \iteme{= 5 :} power-like of the type $\d k_{\perp}^2/(k_{\perp 0}^2 + k_{\perp}^2)$, with $k_{\perp 0}$ in \ttt{PARP(99)} and upper $k_{\perp}$ cut-off given by the $p_{\perp}$ of the hard process or by \ttt{PARP(100)}, whichever is smaller. \itemc{Note:} for options 1 and 2 the \ttt{PARP(100)} value is of minor importance, once \ttt{PARP(100)}$\gg$\ttt{PARP(99)}. However, options 3 and 4 correspond to distributions with infinite $\langle k_{\perp}^2 \rangle$ if the $k_{\perp}$ spectrum is not cut off, and therefore the \ttt{PARP(100)} value is as important for the overall distribution as is \ttt{PARP(99)}. \end{subentry} \iteme{MSTP(94) :} (D=2) (C) energy partitioning in hadron or resolved photon remnant, when this remnant is split into a hadron plus a remainder-jet. The energy fraction chi is taken by one of the two objects, with conventions as described below or for \ttt{PARP(95)} and \ttt{PARP(97)}. \begin{subentry} \iteme{= 1 :} 1 for meson or resolved photon, $2(1-\chi)$ for baryon, i.e. simple counting rules. \iteme{= 2 :} $(k+1)(1-\chi)^k$, with $k$ given by \ttt{PARP(95)} or \ttt{PARP(97)}. \iteme{= 3 :} the $\chi$ of the hadron is selected according to the normal fragmentation function used for the hadron in jet fragmentation, see \ttt{MSTJ(11)}. The possibility of a changed fragmentation function shape in diquark fragmentation (see \ttt{PARJ(45)}) is not included. \iteme{= 4 :} as \ttt{=3}, but the shape is changed as allowed in diquark fragmentation (see \ttt{PARJ(45)}); this change is here also allowed for meson production. (This option is not so natural for mesons, but has been added to provide the same amount of freedom as for baryons). \end{subentry} \iteme{MSTP(131) :}\label{p:MSTP131} (D=0) master switch for pile-up events, i.e. several independent hadron--hadron interactions generated in the same bunch--bunch crossing, with the events following one after the other in the event record. \begin{subentry} \iteme{= 0 :} off, i.e. only one event is generated at a time. \iteme{= 1 :} on, i.e. several events are allowed in the same event record. Information on the processes generated may be found in \ttt{MSTI(41) - MSTI(50)}. \end{subentry} \iteme{MSTP(132) :} (D=4) the processes that are switched on for pile-up events. The first event may be set up completely arbitrarily, using the switches in the \ttt{PYSUBS} common block, while all the subsequent events have to be of one of the `inclusive' processes which dominate the cross section, according to the options below. It is thus not possible to generate two rare events in the pile-up option. \begin{subentry} \iteme{= 1 :} low-$p_{\perp}$ processes (ISUB = 95) only. The low-$p_{\perp}$ model actually used, both in the hard event and in the pile-up events, is the one set by \ttt{MSTP(81)} etc. This means that implicitly also high-$p_{\perp}$ jets can be generated in the pile-up events. \iteme{= 2 :} low-$p_{\perp}$ + double diffractive processes (ISUB = 95 and 94). \iteme{= 3 :} low-$p_{\perp}$ + double diffractive + single diffractive processes (ISUB = 95, 94, 93 and 92). \iteme{= 4 :} low-$p_{\perp}$ + double diffractive + single diffractive + elastic processes, together corresponding to the full hadron--hadron cross section (ISUB = 95, 94, 93, 92 and 91). \end{subentry} \iteme{MSTP(133) :} (D=0) multiplicity distribution of pile-up events. \begin{subentry} \iteme{= 0 :} selected by user, before each \ttt{PYEVNT} call, by giving the \ttt{MSTP(134)} value. \iteme{= 1 :} a Poissonian multiplicity distribution in the total number of pile-up events. This is the relevant distribution if the switches set for the first event in \ttt{PYSUBS} give the same subprocesses as are implied by \ttt{MSTP(132)}. In that case the mean number of events per beam crossing is $\br{n} = \sigma_{\mrm{pile}} \times$\ttt{PARP(31)}, where $\sigma_{\mrm{pile}}$ is the sum of the cross section for allowed processes. Since bunch crossing which do not give any events at all (probability $\exp(-\br{n})$) are not simulated, the actual average number per \ttt{PYEVNT} call is $\langle n \rangle = \br{n}/(1-\exp(-\br{n}))$. \iteme{= 2 :} a biased distribution, as is relevant when one of the events to be generated is assumed to belong to an event class with a cross section much smaller than the total hadronic cross section. If $\sigma_{\mrm{rare}}$ is the cross section for this rare process (or the sum of the cross sections of several rare processes) and $\sigma_{\mrm{pile}}$ the cross section for the processes allowed by \ttt{MSTP(132)}, then define $\br{n} = \sigma_{\mrm{pile}} \times$\ttt{PARP(131)} and $f = \sigma_{\mrm{rare}}/\sigma_{\mrm{pile}}$. The probability that a bunch crossing will give $i$ events is then ${\cal P}_i = f \, i \, \exp(-\br{n}) \, \br{n}^i/i!$, i.e. the na\"{\i}ve Poissonian is suppressed by a factor $f$ since one of the events will be rare rather than frequent, but enhanced by a factor $i$ since any of the $i$ events may be the rare one. Only beam crossings which give at least one event of the required rare type are simulated, and the distribution above normalized accordingly. \itemc{Note:} for practical reasons, it is required that $\br{n} < 120$, i.e. that an average beam crossing does not contain more than 120 pile-up events. The multiplicity distribution is truncated above 200, or when the probability for a multiplicity has fallen below $10^{-6}$, whichever occurs sooner. Also low multiplicities with probabilities below $10^{-6}$ are truncated. See also \ttt{PARI(91) - PARI(93)}. \end{subentry} \iteme{MSTP(134) :} (D=1) a user selected multiplicity, i.e. total number of pile-up events, to be generated in the next \ttt{PYEVNT} call. May be reset for each new event, but must be in the range $1 \leq$\ttt{MSTP(134)}$\leq 200$. \boxsep \iteme{PARP(81) :}\label{p:PARP81} (D=1.40 GeV/$c$) effective minimum transverse momentum $p_{\perp\mathrm{min}}$ for multiple interactions with \ttt{MSTP(82)=1}. \iteme{PARP(82) :} (D=1.55 GeV/$c$) regularization scale $p_{\perp 0}$ of the transverse momentum spectrum for multiple interactions with \ttt{MSTP(82)}$\geq 2$. \iteme{PARP(83), PARP(84) :} (D=0.5, 0.2) parameters of an assumed double Gaussian matter distribution inside the colliding hadrons for \ttt{MSTP(82)=4}, of the form given in eq.~(\ref{mi:doubleGauss}), i.e. with a core of radius \ttt{PARP(84)} of the main radius and containing a fraction \ttt{PARP(83)} of the total hadronic matter. \iteme{PARP(85) :} (D=0.33) probability that an additional interaction in the multiple interaction formalism gives two gluons, with colour connections to `nearest neighbours' in momentum space. \iteme{PARP(86) :} (D=0.66) probability that an additional interaction in the multiple interaction formalism gives two gluons, either as described in \ttt{PARP(85)} or as a closed gluon loop. Remaining fraction is supposed to consist of quark--antiquark pairs. \iteme{PARP(87), PARP(88) :} (D=0.7, 0.5) in order to account for an assumed dominance of valence quarks at low transverse momentum scales, a probability is introduced that a $\mathrm{g}\g$-scattering according to na\"{\i}ve cross section is replaced by a $\mathrm{q}\overline{\mathrm{q}}$ one; this is used only for \ttt{MSTP(82)}$\geq 2$. The probability is parametrized as ${\cal P} = a (1 - (p_{\perp}^2/(p_{\perp}^2 + b^2)^2)$, where $a =$\ttt{PARP(87)} and $b =$\ttt{PARP(88)}$\times$\ttt{PARP(82)}. \iteme{PARP(91) :} (D=0.44 GeV/$c$) (C) width of Gaussian primordial $k_{\perp}$ distribution inside hadron for \ttt{MSTP(91)=1}, i.e. $\exp(-k_{\perp}^2/\sigma^2) \, k_{\perp} \, \d k_{\perp}$ with $\sigma =$\ttt{PARP(91)} and $\langle k_{\perp}^2 \rangle = $\ttt{PARP(91)}$^2$. \iteme{PARP(92) :} (D=0.20 GeV/$c$) (C) width parameter of exponential primordial $k_{\perp}$ distribution inside hadron for \ttt{MSTP(91)=2}, i.e. $\exp(-k_{\perp}/\sigma) \, k_{\perp} \, \d k_{\perp}$ with $\sigma =$\ttt{PARP(92)} and $\langle k_{\perp}^2 \rangle = 6 \times$\ttt{PARP(92)}$^2$. Thus one should put \ttt{PARP(92)}$\approx$\ttt{PARP(91)}$/\sqrt{6}$ to have continuity with the option above. \iteme{PARP(93) :} (D=2. GeV/$c$) (C) upper cut-off for primordial $k_{\perp}$ distribution inside hadron. \iteme{PARP(94) :} (D=1.) (C) for \ttt{MSTP(92)}$\geq 2$ this gives the value of the parameter $k$ for the case when a meson or resolved photon remnant is split into two fragments (which is which is chosen at random). \iteme{PARP(95) :} (D=0.) (C) for \ttt{MSTP(94)=2} this gives the value of the parameter $k$ for the case when a meson or resolved photon remnant is split into a meson and a spectator fragment jet, with $\chi$ giving the energy fraction taken by the meson. \iteme{PARP(96) :} (D=3.) (C) for \ttt{MSTP(92)}$\geq 2$ this gives the value of the parameter $k$ for the case when a nucleon remnant is split into a diquark and a quark fragment, with $\chi$ giving the energy fraction taken by the quark jet. \iteme{PARP(97) :} (D=1.) (C) for \ttt{MSTP(94)=2} this gives the value of the parameter $k$ for the case when a nucleon remnant is split into a baryon and a quark jet or a meson and a diquark jet, with $\chi$ giving the energy fraction taken by the quark jet or meson, respectively. \iteme{PARP(98) :} (D=0.75) (C) for \ttt{MSTP(92)=5} this gives the power of an assumed basic $1/\chi^b$ behaviour in the splitting distribution, with $b =$\ttt{PARP(98)}. \iteme{PARP(99) :} (D=0.44 GeV/$c$) (C) width parameter of primordial $k_{\perp}$ distribution inside photon; exact meaning depends on \ttt{MSTP(93)} value chosen (cf. \ttt{PARP(91)} and \ttt{PARP(92)} above). \iteme{PARP(100) :} (D=2. GeV/$c$) (C) upper cut-off for primordial $k_{\perp}$ distribution inside photon. \iteme{PARP(131) :}\label{p:PARP131} (D=0.01 mb$^{-1}$) in the pile-up events scenario, \ttt{PARP(131)} gives the assumed luminosity per bunch--bunch crossing, i.e. if a subprocess has a cross section $\sigma$, the average number of events of this type per bunch--bunch crossing is $\br{n} = \sigma \times$\ttt{PARP(131)}. \ttt{PARP(131)} may be obtained by dividing the integrated luminosity over a given time (1 s, say) by the number of bunch--bunch crossings that this corresponds to. Since the program will not generate more than 200 pile-up events, the initialization procedure will crash if $\br{n}$ is above 120. \end{entry} \clearpage \section{Fragmentation} The main fragmentation option in {\tsc{Jetset/Pythia}} is the Lund string scheme, but independent fragmentation options are also available. These latter options should not be taken too seriously, since we know that independent fragmentation does not provide a consistent alternative, but occasionally one may like to compare string fragmentation with something else. The subsequent four subsections give further details; the first one on flavour selection, which is common to the two approaches, the second on string fragmentation, the third on independent fragmentation, while the fourth and final contains information on a few other minor issues. The Lund fragmentation model is described in \cite{And83}, where all the basic ideas are presented and earlier papers \cite{And79,And80,And82,And82a} summarized. The details given there on how a multiparton jet system is allowed to fragment are out of date, however, and for this one should turn to \cite{Sjo84}. Also the `popcorn' baryon production mechanism is not covered, see \cite{And85}. Reviews of fragmentation models in general may be found in \cite{Sjo88,Sjo89}. \subsection{Flavour Selection} \label{ss:flavoursel} In either string or independent fragmentation, an iterative approach is used to describe the fragmentation process. Given an initial quark $\mathrm{q} = \mathrm{q}_0$, it is assumed that a new $\mathrm{q}_1 \overline{\mathrm{q}}_1$ pair may be created, such that a meson $\mathrm{q}_0 \overline{\mathrm{q}}_1$ is formed, and a $\mathrm{q}_1$ is left behind. This $\mathrm{q}_1$ may at a later stage pair off with a $\overline{\mathrm{q}}_2$, and so on. What need be given is thus the relative probabilities to produce the various possible $\mathrm{q}_i \overline{\mathrm{q}}_i$ pairs, $\u \overline{\mathrm{u}}$, $\d \overline{\mathrm{d}}$, $\mathrm{s} \overline{\mathrm{s}}$, etc., and the relative probilities that a given $\mathrm{q}_{i-1}\overline{\mathrm{q}}_i$ quark pair combination forms a specific meson, e.g. for $\u \overline{\mathrm{u}}$ either $\pi^+$, $\rho^+$ or some higher state. In {\tsc{Jetset}}, it is assumed that the two aspects can be factorized, i.e. that it is possible first to select a $\mathrm{q}_i \overline{\mathrm{q}}_i$ pair, without any reference to allowed physical meson states, and that, once the $\mathrm{q}_{i-1} \overline{\mathrm{q}}_i$ flavour combination is given, it can be assigned to a given meson state with total probability unity. \subsubsection{Quark flavours and transverse momenta} In order to generate the quark--antiquark pairs $\mathrm{q}_i \overline{\mathrm{q}}_i$ which lead to string breakups, the Lund model invokes the idea of quantum mechanical tunnelling, as follows. If the $\mathrm{q}_i$ and $\overline{\mathrm{q}}_i$ have no (common) mass or transverse momentum, the pair can classically be created at one point and then be pulled apart by the field. If the quarks have mass and/or transverse momentum, however, the $\mathrm{q}_i$ and $\overline{\mathrm{q}}_i$ must classically be produced at a certain distance so that the field energy between them can be transformed into the sum of the two transverse masses $m_{\perp}$. Quantum mechanically, the quarks may be created in one point (so as to keep the concept of local flavour conservation) and then tunnel out to the classically allowed region. In terms of a common transverse mass $m_{\perp}$ of the $\mathrm{q}_i$ and the $\overline{\mathrm{q}}_i$, the tunnelling probability is given by \begin{equation} \exp \left( -\frac{\pi m_{\perp}^2}{\kappa} \right) = \exp \left( -\frac{\pi m^2}{\kappa} \right) \exp \left( -\frac{\pi p_{\perp}^2}{\kappa} \right) ~. \end{equation} The factorization of the transverse momentum and the mass terms leads to a flavour-independent Gaussian spectrum for the $p_x$ and $p_y$ components of $\mathrm{q}_i \overline{\mathrm{q}}_i$ pairs. Since the string is assumed to have no transverse excitations, this $p_{\perp}$ is locally compensated between the quark and the antiquark of the pair. The $p_{\perp}$ of a meson $\mathrm{q}_{i-1} \overline{\mathrm{q}}_i$ is given by the vector sum of the $p_{\perp}$:s of the $\mathrm{q}_{i-1}$ and $\overline{\mathrm{q}}_i$ constituents, which implies Gaussians in $p_x$ and $p_y$ with a width $\sqrt{2}$ that of the quarks themselves. The assumption of a Gaussian shape may be a good first approximation, but there remains the possibility of non-Gaussian tails, that can be important in some situations. In a perturbative QCD framework, a hard scattering is associated with gluon radiation, and further contributions to what is na\"{\i}vely called fragmentation $p_{\perp}$ comes from unresolved radiation. This is used as an explanation why the experimental $\left\langle p_{\perp} \right\rangle$ is somewhat higher than obtained with the formula above. The formula also implies a suppression of heavy quark production $u : d : s : c \approx$ \mbox{$1 : 1 : 0.3 : 10^{-11}$}. Charm and heavier quarks are hence not expected to be produced in the soft fragmentation. Since the predicted flavour suppressions are in terms of quark masses, which are notoriously difficult to assign (should it be current algebra, or constituent, or maybe something in between?), the suppression of $\mathrm{s} \overline{\mathrm{s}}$ production is left as a free parameter in the program: $\u \overline{\mathrm{u}}$ : $\d \overline{\mathrm{d}}$ : $\mathrm{s} \overline{\mathrm{s}}$ = 1 : 1 : $\gamma_s$, where by default $\gamma_s = 0.3$. At least qualitatively, the experimental value agrees with theoretical prejudice. There is no production at all of heavier flavours in the fragmentation process, but only as part of the shower evolution. \subsubsection{Meson production} \label{sss:mesonprod} Once the flavours $\mathrm{q}_{i-1}$ and $\overline{\mathrm{q}}_i$ have been selected, a choice is made between the possible multiplets. The relative composition of different multiplets is not given from first principles, but must depend on the details of the fragmentation process. To some approximation one would expect a negligible fraction of states with radial excitations or non-vanishing orbital angular momentum. Spin counting arguments would then suggest a 3:1 mixture between the lowest lying vector and pseudoscalar multiplets. Wave function overlap arguments lead to a relative enhancement of the lighter pseudoscalar states, which is more pronounced the larger the mass splitting is \cite{And82a}. In the program, six meson multiplets are included. If the nonrelativistic classification scheme is used, i.e. mesons are assigned a valence quark spin $S$ and an internal orbital angular momentum $L$, with the physical spin $s$ denoted $J$, $\mbf{J} = \mbf{L} + \mbf{S}$, then the multiplets are: \begin{Itemize} \item $L = 0$, $S = 0$, $J = 0$: the ordinary pseudoscalar meson multiplet; \item $L = 0$, $S = 1$, $J = 1$: the ordinary vector meson multiplet; \item $L = 1$, $S = 0$, $J = 1$: an axial vector meson multiplet; \item $L = 1$, $S = 1$, $J = 0$: the scalar meson multiplet; \item $L = 1$, $S = 1$, $J = 1$: another axial vector meson multiplet; and \item $L = 1$, $S = 1$, $J = 2$: the tensor meson multiplet. \end{Itemize} Each multiplet has the full four-generation setup of $8 \times 8$ states included in the program, although many could never actually be produced. Some simplifications have been made; thus there is no mixing included between the two axial vector multiplets. In the program, the spin $S$ is first chosen to be either 0 or 1. This is done according to parametrized relative probabilities, where the probability for spin 1 by default is taken to be 0.5 for a meson consisting only of $\u$ and $\d$ quark, 0.6 for one which contains $\mathrm{s}$ as well, and $0.75$ for quarks with $\c$ or heavier quark, in accordance with the deliberations above. By default, it is assumed that $L = 0$, such that only pseudoscalar and vector mesons are produced. For inclusion of $L = 1$ production, four parameters can be used, one to give the probability that a $S = 0$ state also has $L =1$, the other three for the probability that a $S = 1$ state has $L = 1$ and $J$ either 0, 1, or 2. For the flavour-diagonal meson states $\u \overline{\mathrm{u}}$, $\d \overline{\mathrm{d}}$ and $\mathrm{s} \overline{\mathrm{s}}$, it is also necessary to include mixing into the physical mesons. This is done according to a parametrization, based on the mixing angles given in the Review of Particle Properties \cite{PDG88}. In particular, the default choices correspond to \begin{eqnarray} \eta & = & \frac{1}{2} (\u\overline{\mathrm{u}} + \d\overline{\mathrm{d}}) - \frac{1}{\sqrt{2}} \mathrm{s}\overline{\mathrm{s}} ~; \nonumber \\ \eta' & = & \frac{1}{2} (\u\overline{\mathrm{u}} + \d\overline{\mathrm{d}}) + \frac{1}{\sqrt{2}} \mathrm{s}\overline{\mathrm{s}} ~; \nonumber \\ \omega & = & \frac{1}{\sqrt{2}} (\u\overline{\mathrm{u}} + \d\overline{\mathrm{d}}) \nonumber \\ \phi & = & \mathrm{s}\overline{\mathrm{s}} ~. \end{eqnarray} In the $\pi^0 - \eta - \eta'$ system, no account is therefore taken of the difference in masses, an approximation which seems to lead to an overestimate of $\eta'$ rates \cite{ALE92}. Recently, parameters have been introduced to allow an additional `brute force' suppression of $\eta$ and $\eta'$ states. \subsubsection{Baryon production} Baryon production may, in its simplest form, be obtained by assuming that any flavour $\mathrm{q}_i$ given above could represent either a quark or an antidiquark in a colour triplet state. Then the same basic machinery can be run through as above, supplemented with the probability to produce various diquark pairs. In principle, there is one parameter for each diquark, but if tunnelling is still assumed to give an effective description, mass relations can be used to reduce the effective number of parameters. There are three main ones appearing in the program: \begin{Itemize} \item the relative probability to pick a $\overline{\mathrm{q}}\qbar$ diquark rather than a $\mathrm{q}$; \item the extra suppression associated with a diquark containing a strange quark (over and above the ordinary $\mathrm{s} / \u$ suppression factor $\gamma_s$); and \item the suppression of spin 1 diquarks relative to spin 0 ones (apart from the factor of 3 enhancement of the former based on counting the number of spin states). \end{Itemize} The extra strange diquark suppression factor comes about since what appears in the exponent of the tunnelling formula is $m^2$ and not $m$, so that the diquark and the strange quark suppressions do not factorize. Only two baryon multiplets are included, i.e. there are no $L=1$ excited states. The two multiplets are: \begin{Itemize} \item $S = J = 1/2$: the `octet' multiplet of SU(3) (in the full four-generation scenario in the program 168 states are available); \item $S = J = 3/2$: the `decuplet' multiplet of SU(3) (120 states in the program). \end{Itemize} In contrast to the meson case, different flavour combinations have different numbers of states available: for $\u \u \u$ only $\Delta^{++}$, whereas $\u \d \mathrm{s}$ may become either $\Lambda$, $\Sigma^0$ or $\Sigma^{*0}$. An important constraint is that a baryon is a symmetric state of three quarks, neglecting the colour degree of freedom. When a diquark and a quark are joined to form a baryon, the combination is therefore weighted with the probability that they form a symmetric three-quark state. The program implementation of this principle is to first select a diquark at random, with the strangeness and spin 1 suppression factors above included, but then to accept the selected diquark with a weight proportional to the number of states available for the quark-diquark combination. This means that, were it not for the tunnelling suppression factors, all states in the {\bf SU(6)} (flavour {\bf SU(3)} times spin {\bf SU(2)}) 56-multiplet would become equally populated. Of course also heavier baryons may come from the fragmentation of e.g. $\c$ quark jets, but although the particle classification scheme used in the program is {\bf SU(16)}, i.e. with eight flavours, all possible quark-diquark combinations can be related to {\bf SU(6)} by symmetry arguments. As in the case for mesons, one could imagine an explicit further suppression of the heavier spin 3/2 baryons. We do not expect it to be an important effect, since baryon mass splittings are much smaller than in the meson case. In case of rejection, a new diquark is selected and tested, etc. A corresponding procedure is used for the quark selection when a diquark has already been formed in the previous step. Properly speaking both the quark and the diquark flavour should be chosen anew. This would become a tedious process, since also the hadron produced in the step before would have to be rejected. In practice only the last produced pair, be that the quark or diquark one, is rejected. The error introduced by this is small. A more general framework for baryon production is the `popcorn' one \cite{And85}, in which diquarks as such are never produced, but rather baryons appear from the successive production of several $\mathrm{q}_i \overline{\mathrm{q}}_i$ pairs. The picture is the following. Assume that the original $\mathrm{q}$ is red $r$ and the $\overline{\mathrm{q}}$ is $\br{r}$. Normally a new $\mathrm{q}_1 \overline{\mathrm{q}}_1$ pair produced in the field would also be $r \br{r}$, so that the $\overline{\mathrm{q}}_1$ is pulled towards the $\mathrm{q}$ end and vice versa, and two separate colour-singlet systems $\mathrm{q} \overline{\mathrm{q}}_1$ and $\mathrm{q}_1 \overline{\mathrm{q}}$ are formed. Occasionally, the $\mathrm{q}_1 \overline{\mathrm{q}}_1$ pair may be e.g. $g \br{g}$ ($g$ = green), in which case there is no net colour charge acting on either $\mathrm{q}_1$ or $\overline{\mathrm{q}}_1$. Therefore, the pair cannot gain energy from the field, and normally would exist only as a fluctuation. If $\mathrm{q}_1$ moves towards $\mathrm{q}$ and $\overline{\mathrm{q}}_1$ towards $\overline{\mathrm{q}}$, the net field remaining between $\mathrm{q}_1$ and $\overline{\mathrm{q}}_1$ is $\br{b} b$ ($b$ = blue; $g + r = \br{b}$ if only colour triplets are assumed). In this central field, an additional $\mathrm{q}_2 \overline{\mathrm{q}}_2$ pair can be created, where $\mathrm{q}_2$ now is pulled towards $\mathrm{q} \mathrm{q}_1$ and $\overline{\mathrm{q}}_2$ towards $\overline{\mathrm{q}} \overline{\mathrm{q}}_1$, with no net colour field between $\mathrm{q}_2$ and $\overline{\mathrm{q}}_2$. If this is all that happens, the baryon $B$ will be made up out of $\mathrm{q}_1$, $\mathrm{q}_2$ and some $\mathrm{q}_4$ produced between $\mathrm{q}$ and $\mathrm{q}_1$, and $\br{B}$ of $\overline{\mathrm{q}}_1$, $\overline{\mathrm{q}}_2$ and some $\overline{\mathrm{q}}_5$, i.e. the $B$ and $\br{B}$ will be nearest neighbours in rank and share two quark pairs. Specifically, $\mathrm{q}_1$ will gain energy from $\mathrm{q}_2$ in order to end up on mass shell, and the tunnelling formula for an effective $\mathrm{q}_1 \mathrm{q}_2$ diquark is recovered. Part of the time, several $b \br{b}$ colour pair productions may take place between the $\mathrm{q}_1$ and $\overline{\mathrm{q}}_1$, however. With two production vertices $\mathrm{q}_2 \overline{\mathrm{q}}_2$ and $\mathrm{q}_3 \overline{\mathrm{q}}_3$, a central meson $\overline{\mathrm{q}}_2 \mathrm{q}_3$ may be formed, surrounded by a baryon $\mathrm{q}_4 \mathrm{q}_1 \mathrm{q}_2$ and an antibaryon $\overline{\mathrm{q}}_3 \overline{\mathrm{q}}_1 \overline{\mathrm{q}}_5$. We call this a $BM\br{B}$ configuration to distinguish it from the $\mathrm{q}_4 \mathrm{q}_1 \mathrm{q}_2$ + $\overline{\mathrm{q}}_2 \overline{\mathrm{q}}_1 \overline{\mathrm{q}}_5$ $B\br{B}$ configuration above. For $BM\br{B}$ the $B$ and $\br{B}$ only share one quark--antiquark pair, as opposed to two for $B\br{B}$ configurations. The relative probability for a $BM\br{B}$ configuration is given by the uncertainty relation suppression for having the $\mathrm{q}_1$ and $\overline{\mathrm{q}}_1$ sufficiently far apart that a meson may be formed in between. Strictly speaking, also configurations like $BMM\br{B}$, $BMMM\br{B}$, etc. should be possible, but the probability for this is small in our model. Further, since larger masses corresponds to longer string pieces, the production of pseudoscalar mesons is favoured over that of vector ones. If only $B\br{B}$ and $BM\br{B}$ states are included, and if the probability for having a vector meson $M$ is not suppressed extra, two partly compensating errors are made (since a vector meson typically decays into two or more pseudoscalar ones). In total, the flavour iteration procedure therefore contains the following possible subprocesses (plus, of course, their charge conjugates): \begin{Itemize} \item $\mathrm{q}_1 \to \mathrm{q}_2 + (\mathrm{q}_1 \overline{\mathrm{q}}_2)$ meson; \item $\mathrm{q}_1 \to \overline{\mathrm{q}}_2\overline{\mathrm{q}}_3 + (\mathrm{q}_1 \mathrm{q}_2 \mathrm{q}_3)$ baryon; \item $\mathrm{q}_1 \mathrm{q}_2 \to \overline{\mathrm{q}}_3 + (\mathrm{q}_1 \mathrm{q}_2 \mathrm{q}_3)$ baryon; \item $\mathrm{q}_1 \mathrm{q}_2 \to \mathrm{q}_1 \mathrm{q}_3 + (\mathrm{q}_2 \overline{\mathrm{q}}_3)$ meson; \end{Itemize} with the constraint that the last process cannot be iterated to obtain several mesons in between the baryon and the antibaryon. Unfortunately, the resulting baryon production model has a fair number of parameters, which would be given by the model only if quark and diquark masses were known unambiguously. We have already mentioned the $\mathrm{s} / \u$ ratio and the $\mathrm{q}\q / \mathrm{q}$ one; the latter has to be increased from 0.09 to 0.10 for the popcorn model, since the total number of possible baryon production configurations is lower in this case (the particle produced between the $B$ and $\br{B}$ is constrained to be a meson). For the popcorn model, exactly the same parameters as already found in the diquark model are needed to describe the $B\br{B}$ configurations. For $BM\br{B}$ configurations, the square root of a suppression factor should be applied if the factor is relevant only for one of the $B$ and $\br{B}$, e.g. if the $B$ is formed with a spin 1 `diquark' $\mathrm{q}_1 \mathrm{q}_2$ but the $\br{B}$ with a spin 0 diquark $\overline{\mathrm{q}}_1 \overline{\mathrm{q}}_3$. Additional parameters include the relative probability for $BM\br{B}$ configurations, which is assumed to be roughly 0.5 (with the remaining 0.5 being $B\br{B}$), a suppression factor for having a strange meson $M$ between the $B$ and $\br{B}$ (as opposed to having a lighter nonstrange one) and a suppression factor for having a $\mathrm{s} \overline{\mathrm{s}}$ pair (rather than a $\u \overline{\mathrm{u}}$ one) shared between the $B$ and $\br{B}$ of a $BM\br{B}$ configuration. The default parameter values are based on a combination of experimental observation and internal model predictions. In the diquark model, a diquark is expected to have exactly the same transverse momentum distribution as a quark. For $BM\br{B}$ configurations the situation is somewhat more unclear, but we have checked that various possibilities give very similar results. The option implemented in the program is to assume no transverse momentum at all for the $\mathrm{q}_1 \overline{\mathrm{q}}_1$ pair shared by the $B$ and $\br{B}$, with all other pairs having the standard Gaussian spectrum with local momentum conservation. This means that the $B$ and $\br{B}$ $p_{\perp}$:s are uncorrelated in a $BM\br{B}$ configuration and (partially) anticorrelated in the $B\br{B}$ configurations, with the same mean transverse momentum for primary baryons as for primary mesons. Occasionally, the endpoint of a string is not a single parton, but a diquark or antidiquark, e.g. when a quark has been kicked out of a proton beam particle. One could consider fairly complex schemes for the resulting fragmentation. One such \cite{And81} was available in {\tsc{Jetset}} version 6 but is no longer found in version 7. Instead the same basic scheme is used as for diquark pair production above. Thus a $\mathrm{q}\q$ diquark endpoint is let to fragment just as would a $\mathrm{q}\q$ produced in the field behind a matching $\overline{\mathrm{q}}\qbar$ flavour, i.e. either the two quarks of the diquark enter into the same leading baryon, or else a meson is first produced, containing one of the quarks, while the other is contained in the baryon produced in the next step. \subsection{String Fragmentation} An iterative procedure can also be used for other aspects of the fragmentation. This is possible because, in the string picture, the various points where the string break by the production of $\mathrm{q} \overline{\mathrm{q}}$ pairs are causally disconnected. Whereas the space--time picture in the c.m. frame is such that slow particles (in the middle of the system) are formed first, this ordering is Lorentz frame dependent and hence irrelevant. One may therefore make the convenient choice of starting an iteration process at the ends of the string and proceeding towards the middle. The string fragmentation scheme is rather complicated for a generic multiparton state. In order to simplify the discussion, we will therefore start with the simple $\mathrm{q} \overline{\mathrm{q}}$ process, and only later survey the complications that appear when additional gluons are present. (This distinction is made for pedagogical reasons, in the program there is only one general-purpose algorithm). \subsubsection{Fragmentation functions} Assume a $\mathrm{q} \overline{\mathrm{q}}$ jet system, in its c.m. frame, with the quark moving out in the $+z$ direction and the antiquark in the $-z$ one. We have discussed how it is possible to start the flavour iteration from the $\mathrm{q}$ end, i.e. pick a $\mathrm{q}_1 \overline{\mathrm{q}}_1$ pair, form a hadron $\mathrm{q} \overline{\mathrm{q}}_1$, etc. It has also been noted that the tunnelling mechanism is assumed to give a transverse momentum $p_{\perp}$ for each new $\mathrm{q}_i\overline{\mathrm{q}}_i$ pair created, with the $p_{\perp}$ locally compensated between the $\mathrm{q}_i$ and the $\overline{\mathrm{q}}_i$ member of the pair, and with a Gaussian distribution in $p_x$ and $p_y$ separately. In the program, this is regulated by one parameter, which gives the root-mean-square $p_{\perp}$ of a quark. Hadron transverse momenta are obtained as the sum of $p_{\perp}$:s of the constituent $\mathrm{q}_i$ and $\overline{\mathrm{q}}_{i+1}$, where a diquark is considered just as a single quark. What remains to be determined is the energy and longitudinal momentum of the hadron. In fact, only one variable can be selected independently, since the momentum of the hadron is constrained by the already determined hadron transverse mass $m_{\perp}$, \begin{equation} (E+p_z)(E-p_z) = E^2 - p_z^2 = m_{\perp}^2 = m^2 + p_x^2 + p_y^2 ~. \label{fr:massconstr} \end{equation} In an iteration from the quark end, one is led (by the desire for longitudinal boost invariance and other considerations) to select the $z$ variable as the fraction of $E+p_z$ taken by the hadron, out of the available $E+p_z$. As hadrons are split off, the $E+p_z$ (and $E-p_z$) left for subsequent steps is reduced accordingly: \begin{eqnarray} (E+p_z)_{\mrm{new}} & = & (1-z) (E+p_z)_{\mrm{old}} ~, \nonumber \\ (E-p_z)_{\mrm{new}} & = & (E-p_z)_{\mrm{old}} - \frac{m_{\perp}^2}{z (E+p_z)_{\mrm{old}}} ~. \end{eqnarray} The fragmentation function $f(z)$, which expresses the probability that a given $z$ is picked, could in principle be arbitrary --- indeed, several such choices can be used inside the program, see below. If one, in addition, requires that the fragmentation process as a whole should look the same, irrespectively of whether the iterative procedure is performed from the $\mathrm{q}$ end or the $\overline{\mathrm{q}}$ one, `left--right symmetry', the choice is essentially unique \cite{And83a}: the `Lund symmetric fragmentation function', \begin{equation} f(z) \propto \frac{1}{z} z^{a_{\alpha}} \left( \frac{1-z}{z} \right)^{a_{\beta}} \exp \left( - \frac{bm_{\perp}^2}{z} \right) ~. \label{fr:LSFFlong} \end{equation} There is one separate parameter $a$ for each flavour, with the index $\alpha$ corresponding to the `old' flavour in the iteration process, and $\beta$ to the `new' flavour. It is customary to put all $a_{\alpha,\beta}$ the same, and thus arrive at the simplified expression \begin{equation} f(z) \propto z^{-1} (1-z)^a \exp (-bm_{\perp}^2/z) ~. \label{fr:LSFF} \end{equation} In the program, only two separate $a$ values can be given, that for quark pair production and that for diquark one; by default the two are taken to be the same. In addition, there is the $b$ parameter, which is universal. It should be noted that the explicit mass dependence in $f(z)$ implies a harder fragmentation function for heavier hadrons; the asymptotic behaviour of the mean $z$ value for heavy hadrons is \begin{equation} \langle z \rangle \approx 1 - \frac{1+a}{bm_{\perp}^2} ~. \end{equation} Unfortunately it seems this predicts a somewhat harder spectrum for $\mathrm{B}$ mesons than observed in data. For future reference we note that the derivation of $f(z)$ as a by-product also gives the probability distribution in proper time $\tau$ of $\mathrm{q}_i \overline{\mathrm{q}}_i$ breakup vertices. In terms of $\Gamma = (\kappa \tau)^2$, this distribution is \begin{equation} {\cal P}(\Gamma) \, \d \Gamma \propto \Gamma^a \, \exp(-b \Gamma) \, \d \Gamma ~, \end{equation} with the same $a$ and $b$ as above. Many different other fragmentation functions have been proposed, and a few are available as options in the program. \begin{Itemize} \item The Field-Feynman parametrization \cite{Fie78}, \begin{equation} f(z) = 1 - a + 3a(1-z)^2 ~, \end{equation} with default value $a = 0.77$, is intended to be used only for ordinary hadrons made out of $\u$, $\d$ and $\mathrm{s}$ quarks. \item Since there are indications that the shape above is too strongly peaked at $z = 0$, instead a shape like \begin{equation} f(z) = (1+c) (1-z)^c \end{equation} may be used. \item Charm and bottom data clearly indicate the need for a harder fragmentation function for heavy flavours. The best known of these is the Peterson et al.~formula \cite{Pet83} \begin{equation} f(z) \propto \frac{1}{ z \left( 1 - \frac{\displaystyle 1}{\displaystyle z} - \frac{\displaystyle \epsilon_Q}{\displaystyle 1-z} \right)^2 } ~, \label{fr:PetHF} \end{equation} where $\epsilon_Q$ is a free parameter, expected to scale between flavours like $\epsilon_Q \propto 1/m_Q^2$. \item As a crude alternative, that is also peaked at $z=1$, one may use \begin{equation} f(z) = (1+c) z^c ~. \end{equation} \item Bowler \cite{Bow81} has shown, within the framework of the Artru--Mennessier model \cite{Art74}, that a massive endpoint quark with mass $m_Q$ leads to a modification of the symmetric fragmentation function, due to the fact that the string area swept out is reduced for massive endpoint quarks, compared with massless ditto. The Artru--Mennessier model in principle only applies for clusters with a continuous mass spectrum, and does not allow an $a$ term (i.e. $a \equiv 0$); however, it has been shown \cite{Mor89} that, for a discrete mass spectrum, one may still retain an effective $a$ term. In the program an approximate form with an $a$ term has therefore been used: \begin{equation} f(z) \propto \frac{1}{z^{1 + r_Q b m_Q^2}} z^{a_{\alpha}} \left( \frac{1-z}{z} \right)^{a_{\beta}} \exp \left( - \frac{b m_{\perp}^2}{z} \right) ~. \label{fr:LSFFBowler} \end{equation} In principle the prediction is that $r_Q \equiv 1$, but so as to be able to extrapolate smoothly between this form and the Lund symmetric one, it is possible to pick $r_Q$ separately for $\c$, $\b$ and $\t$ hadrons. \end{Itemize} \subsubsection{Joining the jets} The $f(z)$ formula above is only valid, for the breakup of a jet system into a hadron plus a remainder-system, when the remainder mass is large. If the fragmentation algorithm were to be used all the way from the $\mathrm{q}$ end to the $\overline{\mathrm{q}}$ one, the mass of the last hadron to be formed at the $\overline{\mathrm{q}}$ end would be completely constrained by energy and momentum conservation, and could not be on its mass shell. In theory it is known how to take such effects into account, but the resulting formulae are wholly unsuitable for Monte Carlo implementation. The practical solution to this problem is to carry out the fragmentation both from the $\mathrm{q}$ and the $\overline{\mathrm{q}}$ end, such that for each new step in the fragmentation process, a random choice is made as to from what side the step is to be taken. If the step is on the $\mathrm{q}$ side, then $z$ is interpreted as fraction of the remaining $E+p_z$ of the system, while $z$ is interpreted as $E-p_z$ fraction for a step from the $\overline{\mathrm{q}}$ end. At some point, when the remaining mass of the system has dropped below a given value, it is decided that the next breakup will produce two final hadrons, rather than a hadron and a remainder-system. Since the momenta of two hadrons are to be selected, rather than that of one only, there are enough degrees of freedom to have both total energy and total momentum completely conserved. The mass at which the normal fragmentation process is stopped and the final two hadrons formed is not actually a free parameter of the model: it is given by the requirement that the string everywhere looks the same, i.e. that the rapidity spacing of the final two hadrons, internally and with respect to surrounding hadrons, is the same as elsewhere in the fragmentation process. The stopping mass, for a given setup of fragmentation parameters, has therefore been determined in separate runs. If the fragmentation parameters are changed, some retuning should be done but, in practice, reasonable changes can be made without any special arrangements. Consider a fragmentation process which has already split off a number of hadrons from the $\mathrm{q}$ and $\overline{\mathrm{q}}$ sides, leaving behind a a $\mathrm{q}_i \overline{\mathrm{q}}_j$ remainder system. When this system breaks by the production of a $\mathrm{q}_n \overline{\mathrm{q}}_n$ pair, it is decided to make this pair the final one, and produce the last two hadrons $\mathrm{q}_i\overline{\mathrm{q}}_n$ and $\mathrm{q}_n\overline{\mathrm{q}}_j$, if \begin{equation} ( (E+p_z)(E-p_z) )_{\mrm{remaining}} = W_{\mrm{rem}}^2 < W_{\mathrm{min}}^2 ~. \end{equation} The $W_{\mathrm{min}}$ is calculated according to \begin{equation} W_{\mathrm{min}} = ( W_{\mathrm{min} 0} + m_{\mathrm{q} i} + m_{\mathrm{q} j} + k \, m_{\mathrm{q} n} ) \, (1 \pm \delta) ~. \end{equation} Here $W_{\mathrm{min} 0}$ is the main free parameter, typically around 1 GeV, determined to give a flat rapidity plateau (separately for each particle species), while the default $k = 2$ corresponds to the mass of the final pair being taken fully into account. Smaller values may also be considered, depending on what criteria are used to define the `best' joining of the $\mathrm{q}$ and the $\overline{\mathrm{q}}$ chain. The factor $1 \pm \delta$, by default evenly distributed between 0.8 and 1.2, signifies a smearing of the $W_{\mathrm{min}}$ value, to avoid an abrupt and unphysical cut-off in the invariant mass distribution of the final two hadrons. Still, this distribution will be somewhat different from that of any two adjacent hadrons elsewhere. Due to the cut there will be no tail up to very high masses; there are also fewer events close to the lower limit, where the two hadrons are formed at rest with respect to each other. This procedure does not work all that well for heavy flavours, since it does not fully take into account the harder fragmentation function encountered. Therefore, in addition to the check above, one further test is performed for charm and heavier flavours, as follows. If the check above allows more particle production, a heavy hadron $\mathrm{q}_i \overline{\mathrm{q}}_n$ is formed, leaving a remainder $\mathrm{q}_n \overline{\mathrm{q}}_j$. The range of allowed $z$ values, i.e. the fraction of remaining $E+p_z$ that may be taken by the $\mathrm{q}_i \overline{\mathrm{q}}_n$ hadron, is constrained away from 0 and 1 by the $\mathrm{q}_i \overline{\mathrm{q}}_n$ mass and minimal mass of the $\mathrm{q}_n \overline{\mathrm{q}}_j$ system. The limits of the physical $z$ range is obtained when the $\mathrm{q}_n \overline{\mathrm{q}}_j$ system only consists of one single particle, which then has a well-determined transverse mass $m_{\perp}^{(0)}$. From the $z$ value obtained with the infinite-energy fragmentation function formulae, a rescaled $z'$ value between these limits is given by \begin{equation} z' = \frac{1}{2} \left\{ 1 + \frac{m_{\perp i n}^2}{W_{\mrm{rem}}^2} - \frac{m_{\perp n j}^{(0)2}}{W_{\mrm{rem}}^2} + \sqrt{ \left( 1 - \frac{m_{\perp i n}^2}{W_{\mrm{rem}}^2} - \frac{m_{\perp n j}^{(0)2}}{W_{\mrm{rem}}^2} \right)^2 - 4 \frac{m_{\perp i n}^2}{W_{\mrm{rem}}^2} \frac{m_{\perp n j}^{(0)2}}{W_{\mrm{rem}}^2} } \, (2 z -1) \right\} ~. \end{equation} From the $z'$ value, the actual transverse mass $m_{\perp n j} \geq m_{\perp n j}^{(0)}$ of the $\mathrm{q}_n \overline{\mathrm{q}}_j$ system may be calculated. For more than one particle to be produced out of this system, the requirement \begin{equation} m_{\perp n j}^2 = (1-z') \, \left( W_{\mrm{rem}}^2 - \frac{m_{\perp i n}^2}{z'} \right) > (m_{q j} + W_{\mathrm{min} 0})^2 + p_{\perp}^2 \end{equation} has to be fulfilled. If not, the $\mathrm{q}_n \overline{\mathrm{q}}_j$ system is assumed to collapse to one single particle. The consequence of the procedure above is that, the more the infinite energy fragmentation function $f(z)$ is peaked close to $z=1$, the more likely it is that only two particles are produced. In particular, for $\t \overline{\mathrm{t}}$ systems, where very large $\langle z \rangle$ values are predicted, the expectation is that two particle final states will dominate far above the threshold region. The procedure above has been constructed so that the two particle fraction can be calculated directly from the shape of $f(z)$ and the (approximate) mass spectrum, but it is not unique. For the symmetric Lund fragmentation function, a number of alternatives tried all give essentially the same result, whereas other fragmentation functions may be more sensitive to details. Assume now that two final hadrons have been picked. If the transverse mass of the remainder-system is smaller than the sum of transverse masses of the final two hadrons, the whole fragmentation chain is rejected, and started over from the $\mathrm{q}$ and $\overline{\mathrm{q}}$ endpoints. This does not introduce any significant bias, since the decision to reject a fragmentation chain only depends on what happens in the very last step, specifically that the next-to-last step took away too much energy, and not on what happened in the steps before that. If, on the other hand, the remainder-mass is large enough, there are two kinematically allowed solutions for the final two hadrons: the two mirror images in the rest frame of the remainder-system. Also the choice between these two solutions is given by the consistency requirements, and can be derived from studies of infinite energy jets. The probability for the reverse ordering, i.e. where the rapidity and the flavour orderings disagree, is parametrized by \begin{equation} {\cal P}_{\mrm{reverse}} = \frac{1}{2} \left( \frac{m_{\perp i n} + m_{\perp n j}}{W_{\mrm{rem}}} \right)^d ~. \label{fr:revord} \end{equation} For symmetric fragmentation, the ordering is expected to be increasingly strict when the particles involved are more massive. In the program it is therefore assumed that $d$ is a function of the masses, $d = d_0 (m_{\perp i n} + m_{\perp n j})^2$, where $d_0$ is a free parameter. When baryon production is included, some particular problems arise. First consider $B\br{B}$ situations. In the na\"{\i}ve iterative scheme, away from the middle of the event, one already has a quark and is to chose a matching diquark flavour or the other way around. In either case the choice of the new flavour can be done taking into account the number of {\bf SU(6)} states available for the quark-diquark combination. For a case where the final $\mathrm{q}_n \overline{\mathrm{q}}_n$ breakup is an antidiquark-diquark one, the weights for forming $\mathrm{q}_i \overline{\mathrm{q}}_n$ and $\mathrm{q}_n \overline{\mathrm{q}}_i$ enter at the same time, however. We do not know how to handle this problem; what is done is to use weights as usual for the $\mathrm{q}_i \overline{\mathrm{q}}_n$ baryon to select $\mathrm{q}_n$, but then consider $\mathrm{q}_n \overline{\mathrm{q}}_i$ as given (or the other way around with equal probability). If $\mathrm{q}_n \overline{\mathrm{q}}_i$ turns out to be an antidiquark-diquark combination, the whole fragmentation chain is rejected, since we do not know how to form corresponding hadrons. A similar problem arises, and is solved in the same spirit, for a $BM\br{B}$ configuration in which the $B$ (or $\br{B}$) was chosen as third-last particle. When only two particles remain to be generated, it is obviously too late to consider having a $BM\br{B}$ configuration. This is as it should, however, as can be found by looking at all possible ways a hadron of given rank can be a baryon. While some practical compromises have to be accepted in the joining procedure, the fact that the joining takes place in different parts of the string in different events means that, in the end, essentially no visible effects remain. \subsubsection{String motion and infrared stability} We have now discussed the SF scheme for the fragmentation of a simple $\mathrm{q} \overline{\mathrm{q}}$ jet system. In order to understand how these results generalize to arbitrary jet systems, it is first necessary to understand the string motion for the case when no fragmentation takes place. In the following we will assume that quarks as well as gluons are massless, but all arguments can be generalized to massive quarks without too much problem. For a $\mathrm{q} \overline{\mathrm{q}}$ event viewed in the c.m. frame, with total energy $W$, the partons start moving out back-to-back, carrying half the energy each. As they move apart, energy and momentum is lost to the string. When the partons are a distance $W / \kappa$ apart, all the energy is stored in the string. The partons now turn around and come together again with the original momentum vectors reversed. This corresponds to half a period of the full string motion; the second half the process is repeated, mirror-imaged. For further generalizations to multiparton systems, a convenient description of the energy and momentum flow is given in terms of `genes' \cite{Art83}, infinitesimal packets of the four-momentum given up by the partons to the string. Genes with $p_z = E$, emitted from the $\mathrm{q}$ end in the initial stages of the string motion above, will move in the $\overline{\mathrm{q}}$ direction with the speed of light, whereas genes with $p_z = -E$ given up by the $\overline{\mathrm{q}}$ will move in the $\mathrm{q}$ direction. Thus, in this simple case, the direction of motion for a gene is just opposite to that of a free particle with the same four-momentum. This is due to the string tension. If the system is not viewed in the c.m. frame, the rules are that any parton gives up genes with four-momentum proportional to its own four-momentum, but the direction of motion of any gene is given by the momentum direction of the genes it meets, i.e. that were emitted by the parton at the other end of that particular string piece. When the $\mathrm{q}$ has lost all its energy, the $\overline{\mathrm{q}}$ genes, which before could not catch up with $\mathrm{q}$, start impinging on it, and the $\mathrm{q}$ is pulled back, accreting $\overline{\mathrm{q}}$ genes in the process. When the $\mathrm{q}$ and $\overline{\mathrm{q}}$ meet in the origin again, they have completely traded genes with respect to the initial situation. A 3-jet $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ event initially corresponds to having a string piece stretched between $\mathrm{q}$ and $\mathrm{g}$ and another between $\mathrm{g}$ and $\overline{\mathrm{q}}$. Gluon four-momentum genes are thus flowing towards the $\mathrm{q}$ and $\overline{\mathrm{q}}$. Correspondingly, $\mathrm{q}$ and $\overline{\mathrm{q}}$ genes are flowing towards the $\mathrm{g}$. When the gluon has lost all its energy, the $\mathrm{g}$ genes continue moving apart, and instead a third string region is formed in the `middle' of the total string, consisting of overlapping $\mathrm{q}$ and $\overline{\mathrm{q}}$ genes. The two `corners' on the string, separating the three string regions, are not of the gluon-kink type: they do not carry any momentum. If this third region would only appear at a time later than the typical time scale for fragmentation, it could not affect the sharing of energy between different particles. This is true in the limit of high energy, well separated partons. For a small gluon energy, on the other hand, the third string region appears early, and the overall drawing of the string becomes fairly 2-jet-like, since the third string region consists of $\mathrm{q}$ and $\overline{\mathrm{q}}$ genes and therefore behaves exactly as a sting pulled out directly between the $\mathrm{q}$ and $\overline{\mathrm{q}}$. In the limit of vanishing gluon energy, the two initial string regions collapse to naught, and the ordinary 2-jet event is recovered. Also for a collinear gluon, i.e. $\theta_{\mathrm{q} \mathrm{g}}$ (or $\theta_{\overline{\mathrm{q}} \mathrm{g}}$) small, the stretching becomes 2-jet-like. In particular, the $\mathrm{q}$ string endpoint first moves out a distance $\mbf{p}_{\mathrm{q}} / \kappa$ losing genes to the string, and then a further distance $\mbf{p}_{\mathrm{g}} / \kappa$, a first half accreting genes from the $\mathrm{g}$ and the second half re-emitting them. (This latter half actually includes yet another string piece; a corresponding piece appears at the $\overline{\mathrm{q}}$ end, such that half a period of the system involves five different string regions.) The end result is, approximately, that a string is drawn out as if there had only been a single parton with energy $|\mbf{p}_{\mathrm{q}} + \mbf{p}_{\mathrm{g}}|$, such that the simple 2-jet event again is recovered in the limit $\theta_{\mathrm{q} \mathrm{g}} \to 0$. These properties of the string motion are the reason why the string fragmentation scheme is `infrared safe' with respect to soft or collinear gluon emission. The discussions for the 3-jet case can be generalized to the motion of a string with $\mathrm{q}$ and $\overline{\mathrm{q}}$ endpoints and an arbitrary number of intermediate gluons. For $n$ partons, whereof $n-2$ gluons, the original string contains $n-1$ pieces. Anytime one of the original gluons has lost its energy, a new string region is formed, delineated by a pair of `corners'. As the extra `corners' meet each other, old string regions vanish and new are created, so that half a period of the string contains $2n^2 - 6n + 5$ different string regions. Each of these regions can be understood simply as built up from the overlap of (opposite-moving) genes from two of the original partons, according to well specified rules. \subsubsection{Fragmentation of multiparton systems} The full machinery needed for a multiparton system is very complicated, and is described in detail in \cite{Sjo84}. The following outline is far from complete, and is complicated nonetheless. The main message to be conveyed is that a Lorentz covariant algorithm exists for handling an arbitrary parton configuration, but that the necessary machinery is more complex than in either cluster or independent fragmentation. Assume $n$ partons, with ordering along the string, and related four-momenta, given by $\mathrm{q}(p_1) \mathrm{g}(p_2) \mathrm{g}(p_3) \cdots \mathrm{g}(p_{n-1}) \overline{\mathrm{q}}(p_n)$. The initial string then contains $n-1$ separate pieces. The string piece between the quark and its neigbouring gluon is, in four-momentum space, spanned by one side with four-momentum $p_+^{(1)} = p_1$ and another with $p_-^{(1)} = p_2/2$. The factor of 1/2 in the second expression comes from the fact that the gluon shares its energy between two string pieces. The indices `$+$' and `$-$' denotes direction towards the $\mathrm{q}$ and $\overline{\mathrm{q}}$ end, respectively. The next string piece, counted from the quark end, is spanned by $p_+^{(2)} = p_2/2$ and $p_-^{(2)} = p_3/2$, and so on, with the last one being $p_+^{(n-1)} = p_{n-1}/2$ and $p_-^{(n-1)} = p_n$. For the algorithm to work, it is important that all $p_{\pm}^{(i)}$ be light-cone-like, i.e. $p_{\pm}^{(i)2} = 0$. Since gluons are massless, it is only the two endpoint quarks which can cause problems. The procedure here is to create new $p_{\pm}$ vectors for each of the two endpoint regions, defined to be linear combinations of the old $p_{\pm}$ ones for the same region, with coefficients determined so that the new vectors are light-cone-like. De facto, this corresponds to replacing a massive quark at the end of a string piece with a massless quark at the end of a somewhat longer string piece. With the exception of the added fictitious piece, which anyway ends up entirely within the heavy hadron produced from the heavy quark, the string motion remains unchanged by this. In the continued string motion, when new string regions appear as time goes by, the first such string regions that appear can be represented as being spanned by one $p_+^{(j)}$ and another $p_-^{(k)}$ four-vector, with $j$ and $k$ not necessarily adjacent. For instance, in the $\mathrm{q} \mathrm{g} \overline{\mathrm{q}}$ case, the `third' string region is spanned by $p_+^{(1)}$ and $p_-^{(3)}$. Later on in the string evolution history, it is also possible to have regions made up of two $p_+$ or two $p_-$ momenta. These appear when an endpoint quark has lost all its original momentum, has accreted the momentum of an gluon, and is now re-emitting this momentum. In practice, these regions may be neglected. Therefore only pieces made up by a $(p_+^{(j)},p_-^{(k)})$ pair of momenta are considered in the program. The allowes string regions may be ordered in an abstract parameter plane, where the $(j,k)$ indices of the four-momentum pairs define the position of each region along the two (parameter plane) coordinate axes. In this plane the fragmentation procedure can be described as a sequence of steps, starting at the quark end, where $(j,k) = (1,1)$, and ending at the antiquark one, $(j,k) = (n-1,n-1)$. Each step is taken from an `old' $\mathrm{q}_{i-1} \overline{\mathrm{q}}_{i-1}$ pair production vertex, to the production vertex of a `new' $\mathrm{q}_i \overline{\mathrm{q}}_i$ pair, and the string piece between these two string breaks represent a hadron. Some steps may be taken within one and the same region, while others may have one vertex in one region and the other vertex in another region. Consistency requirements, like energy-momentum conservation, dictates that vertex $j$ and $k$ region values be ordered in a monotonic sequence, and that the vertex positions are monotonically ordered inside each region. The four-momentum of each hadron can be read off, for $p_+$ ($p_-$) momenta, by projecting the separation between the old and the new vertex on to the $j$ ($k$) axis. If the four-momentum fraction of $p_{\pm}^{(i)}$ taken by a hadron is denoted $x_{\pm}^{(i)}$, then the total hadron four-momentum is given by \begin{equation} p = \sum_{j=j_1}^{j_2} x_+^{(j)} p_+^{(j)} + \sum_{k=k_1}^{k_2} x_-^{(k)} p_-^{(k)} + p_{x1} \hat{e}_x^{(j_1 k_1)} + p_{y1} \hat{e}_y^{(j_1 k_1)} + p_{x2} \hat{e}_x^{(j_2 k_2)} + p_{y2} \hat{e}_y^{(j_2 k_2)} ~, \label{fr:fourmom} \end{equation} for a step from region $(j_1,k_1)$ to region $(j_2,k_2)$. By necessity, $x_+^{(j)}$ is unity for a $j_1 < j < j_2$, and correspondingly for $x_-^{(k)}$. The $(p_x,p_y)$ pairs are the transverse momenta produced at the two string breaks, and the $(\hat{e}_x,\hat{e}_y)$ pairs four-vectors transverse to the string directions in the regions of the respective string breaks: \begin{eqnarray} & & \hat{e}_x^{(jk)2} = \hat{e}_y^{(jk)2} = -1 ~, \nonumber \\ & & \hat{e}_x^{(jk)} \hat{e}_y^{(jk)} = \hat{e}_{x,y}^{(jk)} p_+^{(j)} = \hat{e}_{x,y}^{(jk)} p_-^{(k)} = 0 ~. \end{eqnarray} The fact that the hadron should be on mass shell, $p^2 = m^2$, puts one constraint on where a new breakup may be, given that the old one is already known, just as eq.~(\ref{fr:massconstr}) did in the simple 2-jet case. The remaining degree of freedom is, as before, to be given by the fragmentation function $f(z)$. The interpretation of the $z$ is only well-defined for a step entirely constrained to one of the initial string regions, however, which is not enough. In the 2-jet case, the $z$ values can be related to the proper times of string breaks, as follows. The variable $\Gamma = (\kappa \tau)^2$, with $\kappa$ the string tension and $\tau$ the proper time between the production vertex of the partons and the breakup point, obeys an iterative relation of the kind \begin{eqnarray} \Gamma_0 & = & 0 ~, \nonumber \\ \Gamma_i & = & (1-z_i) \left( \Gamma_{i-1} + \frac{m_{\perp i}^2}{z_i} \right) ~. \end{eqnarray} Here $\Gamma_0$ represents the value at the $\mathrm{q}$ and $\overline{\mathrm{q}}$ endpoints, and $\Gamma_{i-1}$ and $\Gamma_i$ the values at the old and new breakup vertices needed to produce a hadron with transverse mass $m_{\perp i}$, and with the $z_i$ of the step chosen according to $f(z_i)$. The proper time can be defined in an unambiguous way, also over boundaries between the different string regions, so for multijet events the $z$ variable may be interpreted just as an auxiliary variable needed to determine the next $\Gamma$ value. (In the Lund symmetric fragmentation function derivation, the $\Gamma$ variable actually does appear naturally, so the choice is not as arbitrary as it may seem here.) The mass and $\Gamma$ constraints together are sufficient to determine where the next string breakup is to be chosen, given the preceding one in the iteration scheme. Actually, several ambiguities remain, but are of no importance for the overall picture. The algorithm for finding the next breakup then works something like follows. Pick a hadron, $p_{\perp}$, and $z$, and calculate the next $\Gamma$. If the old breakup is in the region $(j,k)$, and if the new breakup is also assumed to be in the same region, then the $m^2$ and $\Gamma$ constraints can be reformulated in terms of the fractions $x_+^{(j)}$ and $x_-^{(k)}$ the hadron must take of the total four-vectors $p_+^{(j)}$ and $p_-^{(k)}$: \begin{eqnarray} m^2 &=& c_1 + c_2 x_+^{(j)} + c_3 x_-^{(k)} + c_4 x_+^{(j)} x_-^{(k)} ~, \nonumber \\ \Gamma &=& d_1 + d_2 x_+^{(j)} + d_3 x_-^{(k)} + d_4 x_+^{(j)} x_-^{(k)} ~. \label{fr:mGa} \end{eqnarray} Here the coefficients $c_n$ are fairly simple expressions, obtainable by squaring eq.~(\ref{fr:fourmom}), while $d_n$ are slightly more complicated in that they depend on the position of the old string break, but both the $c_n$ and the $d_n$ are explicitly calculable. What remains is an equation system with two unknowns, $x_+^{(j)}$ and $x_-^{(k)}$. The absence of any quadratic terms is due to the fact that all $p_{\pm}^{(i)2} = 0$, i.e. to the choice of a formulation based on light-cone-like longitudinal vectors. Of the two possible solutions to the equation system (elimination of one variable gives a second degree equation in the other), one is unphysical and can be discarded outright. The other solution is checked for whether the $x_{\pm}$ values are actually inside the physically allowed region, i.e. whether the $x_{\pm}$ values of the current step, plus whatever has already been used up in previous steps, are less than unity. If yes, a solution has been found. If no, it is because the breakup could not take place inside the region studied, i.e. because the equation system was solved for the wrong region. One therefore has to change either index $j$ or index $k$ above by one step, i.e. go to the next nearest string region. In this new region, a new equation system of the type in eq.~(\ref{fr:mGa}) may be written down, with new coefficients. A new solution is found and tested, and so on until a physically acceptable solution is found. The hadron four-momentum is now given by an expression of the type (\ref{fr:fourmom}). The breakup found forms the starting point for the new step in the fragmentation chain, and so on. The final joining in the middle is done as in the 2-jet case, with minor extensions. \subsection{Independent Fragmentation} The independent fragmentation (IF) approach dates back to the early seventies \cite{Krz72}, and gained widespread popularity with the Field-Feynman paper \cite{Fie78}. Subsequently, IF was the basis for two programs widely used in the early PETRA/PEP days, the Hoyer et al.~\cite{Hoy79} and the Ali et al.~\cite{Ali80} programs. JETSET has as (non-default) options a wide selection of independent fragmentation algorithms. \subsubsection{Fragmentation of a single jet} In the IF approach, it is assumed that the fragmentation of any system of partons can be described as an incoherent sum of independent fragmentation procedures for each parton separately. The process is to be carried out in the overall c.m. frame of the jet system, with each jet fragmentation axis given by the direction of motion of the corresponding parton in that frame. Exactly as in string fragmentation, an iterative ansatz can be used to describe the sucessive production of one hadron after the next. Assume that a quark is kicked out by some hard interaction, carrying a well-defined amount of energy and momentum. This quark jet $\mathrm{q}$ is split into a hadron $\mathrm{q} \overline{\mathrm{q}}_1$ and a remainder-jet $\mathrm{q}_1$, essentially collinear with each other. New quark and hadron flavours are picked as already described. The sharing of energy and momentum is given by some probability distribution $f(z)$, where $z$ is the fraction taken by the hadron, leaving $1-z$ for the remainder-jet. The remainder-jet is assumed to be just a scaled-down version of the original jet, in an average sense. The process of splitting off a hadron can therefore be iterated, to yield a sequence of hadrons. In particular, the function $f(z)$ is assumed to be the same at each step, i.e. independent of remaining energy. If $z$ is interpreted as the fraction of the jet $E+p_{\mrm{L}}$, i.e. energy plus longitudinal momentum with respect to the jet axis, this leads to a flat central rapidity plateau $dn/dy$ for a large initial energy. Fragmentation functions can be chosen among those listed above for string fragmentation, but also here the default is the Lund symmetric fragmentation function. The normal $z$ interpretation means that a choice of a $z$ value close to $0$ corresponds to a particle moving backwards, i.e. with $p_{\mrm{L}} < 0$. It makes sense to allow only the production of particles with $p_{\mrm{L}} > 0$, but to explicitly constrain $z$ accordingly would destroy longitudinal invariance. The most straightforward way out is to allow all $z$ values but discard hadrons with $p_{\mrm{L}} < 0$. Flavour, transverse momentum and $E + p_{\mrm{L}}$ carried by these hadrons are `lost' for the forward jet. The average energy of the final jet comes out roughly right this way, with a spread of 1--2 GeV around the mean. The jet longitudinal momentum is decreased, however, since the jet acquires an effective mass during the fragmentation procedure. For a 2-jet event this is as it should be, at least on average, because also the momentum of the compensating opposite-side parton is decreased. In addition to local flavour conservation in $\mathrm{q}_i \overline{\mathrm{q}}_i$ splittings, it is also assumed that transverse momentum is locally conserved, i.e. the net $p_{\perp}$ of the $\mathrm{q}_i \overline{\mathrm{q}}_i$ pair as a whole is assumed to be vanishing. The $p_{\perp}$ of the $\mathrm{q}$ is taken to be a Gaussian in the two transverse degrees of freedom separately, with the transverse momentum of a hadron obtained by the sum of constituent quark transverse momenta. Within the IF framework, there is no unique recipe for how gluon jet fragmentation should be handled. One possibility is to treat it exactly like a quark jet, with the initial quark flavour chosen at random among $\u$, $\overline{\mathrm{u}}$, $\d$, $\overline{\mathrm{d}}$, $\mathrm{s}$ and $\overline{\mathrm{s}}$, including the ordinary $\mathrm{s}$ quark suppression factor. Since the gluon is supposed to fragment more softly than a quark jet, the fragmentation fuction may be chosen independently. Another common option is to split the $\mathrm{g}$ jet into a pair of parallel $\mathrm{q}$ and $\overline{\mathrm{q}}$ ones, sharing the energy, e.g. as in a perturbative branching $\mathrm{g} \to \mathrm{q} \overline{\mathrm{q}}$, i.e. $f(z) \propto z^2 + (1-z)^2$. The fragmentation function could still be chosen independently, if so desired. Further, in either case the fragmentation $p_{\perp}$ could be chosen to have a different mean. \subsubsection{Fragmentation of a jet system} In a system of many jets, each jet is fragmented independently. Since each jet by itself does not conserves the flavour, energy and momentum, then neither does a system of jets. At the end of the generation, special algorithms are therefore used to patch this up. The choice of approach has major consequences, e.g. for event shapes and $\alpha_{\mathrm{s}}$ determinations \cite{Sjo84a}. Little attention is usually given to flavour conservation, and we only offer one scheme. When the fragmentation of all jets has been performed, independently of each other, the net initial flavour composition, i.e. number of $\u$ quarks minus number of $\overline{\mathrm{u}}$ quarks etc., is compared with the net final flavour composition. In case of an imbalance, the flavours of the hadron with lowest three-momentum are removed, and the imbalance is re-evaluated. If the remaining imbalance could be compensated by a suitable choice of new flavours for this hadron, flavours are so chosen, a new mass is found and the new energy can be evaluated, keeping the three-momentum of the original hadron. If the removal of flavours from the hadron with lowest momentum is not enough, flavours are removed from the one with next-lowest momentum, and so on until enough freedom is obtained, whereafter the necessary flavours are recombined at random to form the new hadrons. Occasionally one extra $\mathrm{q}_i \overline{\mathrm{q}}_i$ pair must be created, which is then done according to the customary probabilities. Several different schemes for energy and momentum conservation have been devised. One \cite{Hoy79} is to conserve transverse momentum locally within each jet, so that the final momentum vector of a jet is always parallel with that of the corresponding parton. Then longitudinal momenta may be rescaled separately for particles within each jet, such that the ratio of rescaled jet momentum to initial parton momentum is the same in all jets. Since the initial partons had net vanishing three-momentum, so do now the hadrons. The rescaling factors may be chosen such that also energy comes out right. Another common approach \cite{Ali80} is to boost the event to the frame where the total hadronic momentum is vanishing. After that, energy conservation can be obtained by rescaling all particle three-momenta by a common factor. The number of possible schemes is infinite. Two further options are available in the program. One is to shift all particle three-momenta by a common amount to give net vanishing momentum, and then rescale as before. Another is to shift all particle three-momenta, for each particle by an amount proportional to the longitudinal mass with respect to the imbalance direction, and with overall magnitude selected to give momentum conservation, and then rescale as before. In addition, there is a choice of whether to treat separate colour singlets (like $\mathrm{q} \overline{\mathrm{q}}'$ and $\mathrm{q}' \overline{\mathrm{q}}$ in a $\mathrm{q} \overline{\mathrm{q}} \mathrm{q}' \overline{\mathrm{q}}'$ event) separately or as one single big system. A serious conceptual weakness of the IF framework is the issue of Lorentz invariance. The outcome of the fragmentation procedure depends on the coordinate frame chosen, a problem circumvented by requiring fragmentation always to be carried out in the c.m. frame. This is a consistent procedure for 2-jet events, but only a technical trick for multijets. It should be noted, however, that a Lorentz covariant generalization of the independent fragmentation model exists, in which separate `gluon-type' and `quark-type' strings are used, the Montvay scheme \cite{Mon79}. The `quark string' is characterized by the ordinary string constant $\kappa$, whereas a `gluon string' is taken to have a string constant $\kappa_{\mathrm{g}}$. If $\kappa_{\mathrm{g}} > 2 \kappa$ it is always energetically favourable to split a gluon string into two quark ones, and the ordinary Lund string model is recovered. Otherwise, for a 3-jet $\mathrm{q} \overline{\mathrm{q}} \mathrm{g}$ event the three different string pieces are joined at a junction. The motion of this junction is given by the composant of string tensions acting on it. In particular, it is always possible to boost an event to a frame where this junction is at rest. In this frame, much of the standard na\"{\i}ve IF picture holds for the fragmentation of the three jets; additionally, a correct treatment would automatically give flavour, momentum and energy conservation. Unfortunately, the simplicity is lost when studying events with several gluon jets. In general, each event will contain a number of different junctions, resulting in a polypod shape with a number of quark and gluons strings sticking out from a skeleton of gluon strings. With the shift of emphasis from three-parton to multi-parton configurations, the simple option existing in {\tsc{Jetset}}~6.3 therefore is no longer included. A second conceptual weakness of IF is the issue of collinear divergences. In a parton-shower picture, where a quark or gluon is expected to branch into several reasonably collimated partons, the independent fragmentation of one single parton or of a bunch of collinear ones gives quite different outcomes, e.g. with a much larger hadron multiplicity in the latter case. It is conceivable that a different set of fragmentation functions could be constructed in the shower case in order to circumvent this problem (local parton--hadron duality \cite{Dok89} would correspond to having $f(z) = \delta(z-1)$). \subsection{Other Fragmentation Aspects} Here two aspects are considered, which are applicable regardless of whether string or independent fragmentation is used. \subsubsection{Small mass systems} Occasionally, a jet system may have too small an invariant mass for the ordinary jet fragmentation schemes. This is particularly a problem when showers are used, since two nearby $\mathrm{g} \to \mathrm{q}' \overline{\mathrm{q}}'$ branchings may give rise to an intermediate low-mass colour-singlet system. Before the ordinary fragmentation, one includes an optional additional step, to catch situations of this kind. First the jet system with lowest invariant mass, minus endpoint quark masses, is found. If this is too low for jet fragmentation, an attempt is made to split the system into two hadrons by producing a new $\mathrm{q}_n \overline{\mathrm{q}}_n$ pair (with $\mathrm{q}_n$ chosen according to the standard fragmentation scheme, so that e.g. also diquarks are allowed) to go with the existing endpoint flavours. If the sum of the two thus constructed hadron masses is smaller than the total invariant mass, a simple isotropic two-particle decay is performed. If not, the endpoint flavours are combined to give one single hadron. Next, the parton (or hadron) is found which, when taken together with the jet system, has the largest invariant mass. A minimal transfer of four-momentum is then performed, which puts the hadron on mass shell while keeping the mass of the parton unchanged. With this done, one may again search for a low-mass jet system, and iterate the procedure above, if need be. The procedure may be seen as a `poor man's cluster fragmentation', i.e. a cluster and a low-mass string are considered to be more or less the same thing. \subsubsection{Bose--Einstein effects} A crude option for the simulation of Bose--Einstein effects is included, but is turned off by default. Here the detailed physics is not that well understood, see e.g. \cite{Lor89}. What is offered is an algorithm, more than just a parametrization (since very specific assumptions and choices have been made), and yet less than a true model (since the underlying physics picture is rather fuzzy). In this scheme, the fragmentation is allowed to proceed as usual, and so is the decay of short-lived particles like $\rho$. Then pairs of identical particles, $\pi^+$ say, are considered one by one. The $Q_{ij}$ value of a pair $i$ and $j$ is evaluated, \begin{equation} Q_{ij} = \sqrt{ (p_i + p_j)^2 - 4m^2} ~, \end{equation} where $m$ is the common particle mass. A shifted (smaller) $Q'_{ij}$ is then to be found such that the (infinite statistics) ratio $C_2(Q)$ of shifted to unshifted $Q$ distributions is given by the requested parametrization. The shape may be chosen either exponential or Gaussian, \begin{equation} C_2(Q) = 1 + \lambda \exp \left( - (Q/d)^r \right), ~~~~r = 1~\mrm{or}~2 ~. \end{equation} (In fact, the distribution has to dip slightly below unity at $Q$ values outside the Bose enhancement region, from conservation of total multiplicity.) If the inclusive distribution of $Q_{ij}$ values is assumed given just by phase space, at least at small relative momentum then, with $\d^3 p / E \propto Q^2 \, \d Q / \sqrt{Q^2 + 4m^2}$, then $Q'_{ij}$ is found as the solution to the equation \begin{equation} \int_0^{Q_{ij}} \frac{Q^2 \, \d Q}{\sqrt{Q^2 + 4 m^2}} = \int_0^{Q'_{ij}} C_2(Q) \, \frac{Q^2 \, \d Q}{\sqrt{Q^2 + 4 m^2}} ~. \end{equation} The change of $Q_{ij}$ can be translated into an effective shift of the three-momenta of the two particles, if one uses as extra constraint that the total three-momentum of each pair be conserved in the c.m. frame of the event. Only after all pairwise momentum shifts have been evaluated, with respect to the original momenta, are these momenta actually shifted, for each particle by the sum of evaluated shifts. The total energy of the event is slightly reduced in the process, which is compensated by an overall rescaling of all c.m. frame momentum vectors. It can be discussed which are the particles to involve in this rescaling. Currently the only exceptions to using everything are leptons and neutrinos coming from resonance decays (such as $\mathrm{W}$'s) and photons radiated by leptons (also in initial state radiation). Finally, the decay chain is resumed with more long-lived particles like $\pi^0$. Two comments can be made. The Bose--Einstein effect is here interpreted almost as a classical force acting on the `final state', rather than as a quantum mechanical phenomenon on the production amplitude. This is not a credo, but just an ansatz to make things manageable. Also, since only pairwise interactions are considered, the effects associated with three or more nearby particles tend to get overestimated. (More exact, but also more time-consuming methods may be found in \cite{Zaj87}.) Thus the input $\lambda$ may have to be chosen smaller than what one wants to get out. (On the other hand, many of the pairs of an event contains at least one particle produced in some secondary vertex, like a $\mathrm{D}$ decay. This reduces the fraction of pairs which may contribute to the Bose--Einstein effects, and thus reduces the potential signal.) This option should therefore be used with caution, and only as a first approximation to what Bose--Einstein effects can mean. \clearpage \section{Particles and Their Decays} Particles are the building blocks from which events are constructed. We here use the word `particle' in its broadest sense, i.e. including partons, resonances, hadrons, and so on, subgroups we will describe in the following. Each particle is characterized by some quantities, such as charge and mass. In addition, many of the particles are unstable and subsequently decay. This section contains a survey of the particle content of the programs, and the particle properties assumed. In particular, the decay treatment is discussed. Some particle and decay properties form part already of the hard subprocess description, and are therefore described in sections \ref{s:JETSETproc}, \ref{s:PYTprocgen} and \ref{s:pytproc}. \subsection{The Particle Content} \label{ss:decpartcont} In order to describe both current and potential future physics, a number of different particles are needed. A list of some particles, along with their codes, is given in section \ref{ss:codes}. Here we therefore emphasize the generality rather than the details. Four full generations of quarks and leptons are included in the program, although indications from LEP strongly suggest that only three exist in Nature. There is no standard terminology for the fourth generation; we use $\mathrm{l}$ for the down type quark ($\mathrm{l}$ for low), $\mathrm{h}$ for the up type quark ($\mathrm{h}$ for high), $\chi$ for the lepton and $\nu_{\chi}$ for the neutrino. Quarks may appear either singly or in pairs; the latter are called diquarks and are characterized by their flavour content and their spin. A diquark is always assumed to be in a colour antitriplet state. From the coloured quarks (and diquarks), the colour neutral hadrons may be build up. Six full meson multiplets are included and two baryon ones, see section \ref{ss:flavoursel}. In addition, $\mathrm{K}_{\mrm{S}}^0$ and $\mathrm{K}_{\mrm{L}}^0$ are considered as separate particles coming from the `decay' of $\mathrm{K}^0$ and $\br{\mathrm{K}}^0$ (or, occasionally, produced directly). Other particles from the Standard Model include the gluon $\mathrm{g}$, the photon $\gamma$, the intermediate gauge bosons $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$, and the standard Higgs $\H^0$. Non-standard particles include additional gauge bosons, $\mathrm{Z}'^0$ and $\mathrm{W}'^{\pm}$, additional Higgs bosons $\H'^0$, $\mathrm{A}^0$ and $\H^{\pm}$, a leptoquark $\L_{\mathrm{Q}}$ and a horizontal gauge boson $\mathrm{R}^0$. It is also possible to use the particle codes of the current fourth generation fermions to represent excited quarks and leptons, $\mathrm{q}^*$ and $\ell^*$. From the point of view of usage inside the programs, particles may be subdivided into three classes, partly overlapping. \begin{Enumerate} \item A parton is generically any object which may be found in the wave function of the incoming beams, and may participate in initial- or final-state showers. This includes what is normally meant by partons, i.e. quarks and gluons, but here also leptons and photons. In a few cases other particles may be classified as partons in this sense. \item A resonance is an unstable particle produced as part of the hard process, and where the decay treatment normally is also part of the hard process. Resonance partial widths are perturbatively calculable, and therefore it is possible to dynamically recalculate branching ratios as a function of the mass assigned to a resonance. Resonances includes particles like the $\mathrm{Z}^0$ and other massive gauge bosons and Higgs particles. It does not include hadrons with non-vanishing width, like the $\rho$, which are just called `unstable hadrons'. \item Hadrons, i.e. mesons and baryons produced either in the fragmentation process, in secondary decays or as part of the beam remnant treatment, but not directly as part of the hard process (except in a few special cases). Hadrons may be stable or unstable. Branching ratios are not assumed perturbatively calculable, and can therefore be set freely. Also leptons and photons produced in decays belong to this class. \end{Enumerate} Usually the subdivision above is easy to understand and gives you the control you would expect. However, the classification of top and the fourth generation fermions may lead to some confusion, as already mentioned, section \ref{sss:heavflavclass}. The problem is that the top did not use to be treated as a resonance, but was rather allowed to fragment to hadrons, which subsequently decayed. This approach was a reasonable choice in the days when the top mass was assumed quite light by today's standards. However, given current top limits, the fragmentation and the decay of the top quark is being played out on comparable time scales, and the treatment becomes much more difficult (see e.g. ref. \cite{Sjo92a} for a toy model description). Starting at masses of around 120 GeV the top decay time is so short that no top hadrons at all are formed, and then a true resonance description is appropriate, but still with some complications due to the net colour charge of the top quark. Such an option is now default, wherein the top quark is assumed to decay immediately, but intermediate scenarios can not be modelled. The appearance of intermediate top hadrons in the description has little influence on event shapes, even for a very heavy top. \subsection{Masses, Widths and Lifetimes} \subsubsection{Masses} Quark masses are not particularly well defined. In the program it is necessary to make use of two kinds of masses, current algebra ones and constituent ones. The former are relevant for the kinematics in hard processes (e.g. in $\mathrm{g}\g \to \c\overline{\mathrm{c}}$) and for couplings to Higgs particles, and therefore directly affect cross sections. These values are the ones stored in the standard mass array \ttt{PMAS}. Constituent masses are used to derive the masses of hadrons, and are stored separately in the \ttt{PARF} array. We maintain this distinction for the five first flavours, using the following values by default: \\ \begin{tabular}{ccc@{\protect\rule{0mm}{\tablinsep}}} quark & current algebra mass & constituent mass \\ d & 0.0099 GeV & 0.325 GeV \\ u & 0.0056 GeV & 0.325 GeV \\ s & 0.199 GeV & 0.5 GeV \\ c & 1.35 GeV & 1.6 GeV \\ b & 5.0 GeV & 5.0 GeV. \\ \end{tabular} \\ For top and fourth generation quarks the distinction is not as important, so only one set of mass values is used, namely the one in \ttt{PMAS}. The default top mass is 160 GeV. Constituent masses for diquarks are defined as the sum of the respective quark masses. The gluon is always assumed massless. Particle masses, when known, are taken from ref. \cite{PDG92}. Hypothesized particles, such as fourth generation fermions and Higgs bosons, are assigned some not unreasonable set of default values, in the sense of where you want to search for them in the not too distant future. Here it is understood that you will go in and change the default values according to your own opinions at the beginning of a run. The total number of hadrons in the program is very large, whereof many are not yet discovered (like charm and bottom baryons) and other may or may not exist (top and fourth generation hadrons). In particular for the latter, it would be messy for the user to have to recalculate the masses of hadron each time the assumed quark mass is changed. Therefore the masses of yet undiscovered mesons and baryons are built up, when needed, from the constituent masses. For this purpose one uses formulae of the type \cite{DeR75} \begin{equation} m = m_0 + \sum_i m_i + k \, m_{\d}^2 \sum_{i<j} \frac{\langle \mbox{\boldmath $\sigma$}_i \cdot \mbox{\boldmath $\sigma$}_j \rangle}{m_i \, m_j} ~, \end{equation} i.e. one constant term, a sum over constituent masses and a spin-spin interaction term for each quark pair in the hadron. The constants $m_0$ and $k$ are fitted from known masses, treating mesons and baryons separately. For mesons with orbital angular momentum $L=1$ the spin-spin coupling is assumed vanishing, and only $m_0$ is fitted. One may also define `constituent diquarks masses' using the formula above, with a $k$ value $2/3$ that of baryons. The default values are: \\ \begin{tabular}{ccc@{\protect\rule{0mm}{\tablinsep}}} multiplet & $m_0$ & $k$ \\ pseudoscalars and vectors & 0. & 0.16 GeV \\ axial vectors ($S=0$) & 0.50 GeV & 0. \\ scalars & 0.45 GeV & 0. \\ axial vectors ($S=1$) & 0.55 GeV & 0. \\ tensors & 0.60 GeV & 0. \\ baryons & 0.11 GeV & 0.048 GeV \\ diquarks & 0.077 GeV & 0.048 GeV.\\ \end{tabular} \\ There is one exception to the rule above, and that is flavour neutral mesons, i.e. the onia states of a heavy quark--antiquark pair. These are defined individually, to allow more flexibility. \subsubsection{Widths} A width is calculated perturbatively for those resonances which appear in the {\tsc{Pythia}} hard process generation machinery. The width is used to select masses in hard processes according to a relativistic Breit--Wigner shape. In many processes the width is allowed to be $\hat{s}$-dependent, see section \ref{ss:kinemreson}. Other particle masses, as discussed so far, have been fixed at their nominal value, i.e. with no mass broadening for short-lived particles such as $\rho$, $\mathrm{K}^*$ or $\Delta$. Compared to the $\mathrm{Z}^0$, it is much more difficult to describe the $\rho$ resonance shape, since nonperturbative and threshold effects act to distort the na\"{\i}ve shape. Thus the $\rho$ mass is limited from below by its decay $\rho \to \pi\pi$, but also from above, e.g. in the decay $\phi \to \rho \pi$. In some decay chains, several mass choices are coupled, like in $\a_2 \to \rho \pi$, where also the $\a_2$ has a non-negligible width. Finally, there are some extreme cases, like the $\mathrm{f}_0$, which has a nominal mass below the $\mathrm{K}\K$ threshold, but a tail extending beyond that threshold, and therefore a non-negligible branching ratio to the $\mathrm{K}\K$ channel. In view of examples like these, no attempt is made to provide a full description. Instead a simplified description is used, which should be enough to give the general smearing of events due to mass broadening, but maybe not sufficient for detailed studies of a specific resonance. By default, hadrons are therefore given a mass distribution according to a non-relativistic Breit--Wigner \begin{equation} {\cal P}(m) \, \d m \propto \frac{1}{(m - m_0)^2 + \Gamma^2/4} \, \d m ~. \label{dec:BWlin} \end{equation} Leptons and resonances not taken care of by the hard process machinery are distributed according to a relativistic Breit--Wigner \begin{equation} {\cal P}(m^2) \, \d m^2 \propto \frac{1}{(m^2 - m_0^2)^2 + m_0^2 \Gamma^2} \, \d m^2 ~. \label{dec:BWtwo} \end{equation} Here $m_0$ and $\Gamma$ are the nominal mass and width of the particle. The Breit--Wigner shape is truncated symmetrically, $|m - m_0| < \delta$, with $\delta$ arbitrarily chosen for each particle so that no problems are encountered in the decay chains. It is possible to switch off the mass broadening, or to use either a non-relativistic or a relativistic Breit--Wigners everywhere. The $\mathrm{f}_0$ problem has been `solved' by shifting the $\mathrm{f}_0$ mass to be slightly above the $\mathrm{K}\K$ threshold and have vanishing width. Then kinematics in decays $\mathrm{f}_0 \to \mathrm{K}\K$ is reasonably well modelled. The $\mathrm{f}_0$ mass is too large in the $\mathrm{f}_0 \to \pi\pi$ channel, but this does not really matter, since one anyway is far above threshold here. \subsubsection{Lifetimes} Clearly the lifetime and the width of a particle are inversely related. For practical applications, however, any particle with a non-negligible width decays too close to its production vertex for the lifetime to be of any interest. In the program, the two aspects are therefore considered separately. Particles with a non-vanishing nominal proper lifetime $\tau_0 = \langle \tau \rangle$ are assigned an actual lifetime according to \begin{equation} {\cal P}(\tau) \, \d \tau \propto \exp(- \tau / \tau_0 ) \, \d \tau ~, \end{equation} i.e. a simple exponential decay is assumed. Since the program uses dimensions where the speed of light $c \equiv 1$, and space dimensions are in mm, then actually the unit of $c \tau_0$ is mm and of $\tau_0$ itself mm$/c \approx 3.33\times10^{-12}$ s. If a particle is produced at a vertex $v = (\mbf{x}, t)$ with a momentum $p = (\mbf{p}, E)$ and a lifetime $\tau$, the decay vertex position is assumed to be \begin{equation} v' = v + \tau \, \frac{p}{m} ~, \label{dec:newvertex} \end{equation} where $m$ is the mass of the particle. With the primary interaction (normally) in the origin, it is therefore possible to construct all secondary vertices in parallel with the ordinary decay treatment. The formula above does not take into account any detector effects, such as a magnetic field. It is therefore possible to stop the decay chains at some suitable point, and leave any subsequent decay treatment to the detector simulation program. One may select that particles are only allowed to decay if they have a nominal lifetime $\tau_0$ shorter than some given value or, alternatively, if their decay vertices $\mbf{x}'$ are inside some spherical or cylindrical volume around the origin. \subsection{Decays} \label{ss:partdecays} Several different kinds of decay treatment are used in the program, depending on the nature of the decay. Not discussed here are the decays of resonances which are handled as part of the hard process. \subsubsection{Strong and electromagnetic decays} The decays of hadrons containing the `ordinary' $\u$, $\d$ and $\mathrm{s}$ quarks into two or three particles are known, and branching ratios may be found in \cite{PDG92}. We normally assume that the momentum distributions are given by phase space. There are a few exceptions, where the phase space is weighted by a matrix-element expression, as follows. In $\omega$ and $\phi$ decays to $\pi^+ \pi^- \pi^0$, a matrix element of the form \begin{equation} |{\cal M}|^2 \propto | \mbf{p}_{\pi^+} \times \mbf{p}_{\pi^-} |^2 \label{dec:omegphi} \end{equation} is used, with the $\mbf{p}_{\pi}$ the pion momenta in the rest frame of the decay. (Actually, what is coded is the somewhat more lengthy Lorentz invariant form of the expression above.) Consider the decay chain $P_0 \to P_1 + V \to P_1 + P_2 + P_3$, with $P$ representing pseudoscalar mesons and $V$ a vector one. Here the decay angular distribution of the $V$ in its rest frame is \begin{equation} |{\cal M}|^2 \propto \cos^2 \theta_{02} ~, \label{dec:psvpsps} \end{equation} where $\theta_{02}$ is the angle between $P_0$ and $P_2$. The classical example is $\mathrm{D} \to \mathrm{K}^* \pi \to \mathrm{K} \pi \pi$. If the $P_1$ is replaced by a $\gamma$, the angular distribution in the $V$ decay is instead $\propto \sin^2 \theta_{02}$. In Dalitz decays, $\pi^0$ or $\eta \to \mathrm{e}^+\mathrm{e}^- \gamma$, the mass $m^*$ of the $\e^+\e^-$ pair is selected according to \begin{equation} {\cal P}(m^{*2}) \, \d m^{*2} \propto \frac{\d m^{*2}}{m^{*2}} \, \left( 1 + \frac{2m_{\mathrm{e}}^2}{m^{*2}} \right) \, \sqrt{ 1 - \frac{4m_{\mathrm{e}}^2}{m^{*2}} } \, \left( 1 - \frac{m^{*2}}{m_{\pi,\eta}^2} \right)^3 \, \frac{1}{ (m_{\rho}^2 - m^{*2})^2 + m_{\rho}^2 \Gamma_{\rho}^2 } ~. \label{dec:Dalitz} \end{equation} The last factor, the VMD-inspired $\rho^0$ propagator, is negligible for $\pi^0$ decay. Once the $m^*$ has been selected, the angular distribution of the $\e^+\e^-$ pair is given by \begin{equation} |{\cal M}|^2 \propto (m^{*2} - 2 m_{\mathrm{e}}^2) \left\{ (p_{\gamma} p_{\mathrm{e}^+})^2 + (p_{\gamma} p_{\mathrm{e}^-})^2 \right\} + 4m_{\mathrm{e}}^2 \left\{ (p_{\gamma} p_{\mathrm{e}^+}) (p_{\gamma} p_{\mathrm{e}^-}) + (p_{\gamma} p_{\mathrm{e}^+})^2 + (p_{\gamma} p_{\mathrm{e}^-})^2 \right\} ~. \end{equation} Also a number of simple decays involving resonances of heavier hadrons, e.g. $\Sigma_{\c}^0 \to \Lambda_{\c}^+ \pi^-$ or $\mathrm{B}^{*-} \to \mathrm{B}^- \gamma$ are treated in the same way as the other two-particle decays. \subsubsection{Weak decays of charm hadrons} The charm hadrons have a mass in an intermediate range, where the effects of the na\"{\i}ve $V-A$ weak decay matrix element is partly but not fully reflected in the kinematics of final-state particles. Therefore different decay strategies ar combined. We start with hadronic decays, and subseqently consider semileptonic ones. For the four `main' charm hadrons, $\mathrm{D}^+$, $\mathrm{D}^0$, $\mathrm{D}_{\mathrm{s}}^+$ and $\Lambda_{\c}^+$, a number of branching ratios are already known. The known braching ratios have been combined with reasonable guesses, to construct more or less complete tables of all channels. For hadronic decays of $\mathrm{D}^0$ and $\mathrm{D}^+$, where rather much is known, all channels have an explicitly listed particle content. However, only for the two-body decays is resonance production properly taken into account. It means that the experimentally measured branching ratio for a $\mathrm{K} \pi \pi$ decay channel, say, is represented by contributions from a direct $\mathrm{K} \pi \pi$ channel as well as from indirect ones, such as $\mathrm{K}^* \pi$ and $\mathrm{K} \rho$. For a channel like $\mathrm{K} \pi \pi \pi$, on the other hand, only the $\mathrm{K}^* \rho$ appears separately, while the rest is lumped into one entry in the decay tables. This is more or less in agreement with the philosophy adopted in the PDG tables \cite{PDG92}. For $\mathrm{D}_{\mathrm{s}}^+$ and $\Lambda_{\c}^+$ knowledge is rather incomplete, and only two-body decay channels are listed. Final states with three or more hadron are only listed in terms of a flavour content. The way the program works, it is important to include all the allowed decay channels up to a given multiplicity. Channels with multiplicity higher than this may then be generated according to a simple flavour combination scheme. For instance, in a $\mathrm{D}_{\mathrm{s}}^+$ decay, the normal quark content is $\mathrm{s}\overline{\mathrm{s}}\u\overline{\mathrm{d}}$, where one $\overline{\mathrm{s}}$ is the spectator quark and the others come from the weak decay of the $\c$ quark. The spectator quark may also be annihilated, like in $\mathrm{D}_{\mathrm{s}}^+ \to \u\overline{\mathrm{d}}$. The flavour content to make up one or two hadrons is therefore present from the onset. If one decides to generate more hadrons, this means new flavour-antiflavour pairs have to be generated and combined with the existing flavours. This is done using the same flavour approach as in fragmentation. In more detail, the following scheme is used. \begin{Enumerate} \item The multiplicity is first selected. The $\mathrm{D}_{\mathrm{s}}^+$ and $\Lambda_{\c}^+$ multiplicity is selected according to a distribution described further below. The program can also be asked to generate events of a predetermined multiplicity. \item One of the non-spectator flavours is selected at random. This flavour is allowed to `fragment' into a hadron plus a new remaining flavour, using exactly the same flavour generation algorithm as in the standard jet fragmentation, section \ref{ss:flavoursel}. \item Step 2 is iterated until only one or two hadrons remain to be generated, depending on whether the original number of flavours is two or four. In each step one `unpaired' flavour is replaced by another one as a hadron is `peeled off', so the number of unpaired flavours is preserved. \item If there are two flavours, these are combined to form the last hadron. If there are four, then one of the two possible pairings into two final hadrons is selected at random. To find the hadron species, the same flavour rules are used as when final flavours are combined in the joining of two jets. \item If the sum of decay product masses is larger than the mass of the decaying particle, the flavour selection is rejected and the process is started over at step 1. Normally a new multiplicity is picked, but for $\mathrm{D}^0$ and $\mathrm{D}^+$ the old multiplicity is retained. \item Once an acceptable set of hadrons has been found, these are distributed according to phase space. \end{Enumerate} The picture then is one of a number of partons moving apart, fragmenting almost like jets, but with momenta so low that phase-space considerations are enough to give the average behaviour of the momentum distribution. Like in jet fragmentation, endpoint flavours are not likely to recombine with each other. Instead new flavour pairs are created in between them. One should also note that, while vector and pseudoscalar mesons are produced at their ordinary relative rates, events with many vectors are likely to fail in step 5. Effectively, there is therefore a shift towards lighter particles, especially at large multiplicities. When a multiplicity is to be picked, this is done according to a Gaussian distribution, centered at $c + n_{\mathrm{q}}/4$ and with a width $\sqrt{c}$, with the final number rounded off to the nearest integer. The value for the number of quarks $n_{\mathrm{q}}$ is 2 or 4, as described above, and \begin{equation} c = c_1 \, \ln \left( \frac{m - \sum m_{\mathrm{q}}}{c_2} \right) ~, \label{dec:multsel} \end{equation} where $m$ is the hadron mass and $c_1$ and $c_2$ have been tuned to give a reasonable description of multiplicities. There is always some lower limit for the allowed multiplicity; if a number smaller than this is picked the choice is repeated. Since two-body decays are explicitly enumerated for $\mathrm{D}_{\mathrm{s}}^+$ and $\Lambda_{\c}^+$, there the mimimum multiplicity is three. Semileptonic branching ratios are explicitly given in the program for all the four particles discussed here, i.e. it is never necessary to generate the flavour content using the fragmentation description. This does not mean that all branching ratios are known; a fair amount of guesswork is involved for the channels with higher multiplicities, based on a knowledge of the inclusive semileptonic branching ratio and the exclusive branching ratios for low multiplicities. In semileptonic decays it is not appropriate to distribute the lepton and neutrino momenta according to phase space. Instead the simple $V-A$ matrix element is used, in the limit that decay product masses may be neglected and that quark momenta can be replaced by hadron momenta. Specifically, in the decay $H \to \ell^+ \nu_{\ell} h$, where $H$ is a charm hadron and $h$ and ordinary hadron, the matrix element \begin{equation} |{\cal M}|^2 = (p_H p_{\ell}) (p_{\nu} p_h) \end{equation} is used to distribute the products. It is not clear how to generalize this formula when several hadrons are present in the final state. In the program, the same matrix element is used as above, with $p_h$ replaced by the total four-momentum of all the hadrons. This tends to favour a low invariant mass for the hadronic system compared with na\"{\i}ve phase space. There are a few charm hadrons, such as $\Xi_c$ and $\Omega_c$, which decay weakly but are so rare that little is known about them. For these a simplified generic charm decay treatment is used. For hadronic decays only the quark content is given, and then a multiplicity and a flavour composition is picked at random, as already described. Semileptonic decays are assumed to produce only one hadron, so that $V-A$ matrix element can be simply applied. \subsubsection{Weak decays of the $\tau$ lepton} For the $\tau$ lepton, an explicit list of decay channels has been put together, which includes channels with up to five final-state particles, some of which may be unstable and subsequently decay to produce even larger total multiplicities. Because of the well-known `$\tau$ puzzle', i.e. that experimentally the sum of branching ratios for exclusive one-prong decays is lower than the inclusive one-prong branching ratio, such a table cannot be constructed in full agreement with the PDG data. (The problem is nowadays less severe than it used to be, but still not fully resolved.) The leptonic decays $\tau^- \to \nu_{\tau} \ell^- \br{\nu}_{\ell}$, where $\ell$ is $\mathrm{e}$ or $\mu$, are distributed according to the standard $V-A$ matrix element \begin{equation} |{\cal M}|^2 = (p_{\tau} p_{\br{\nu}_{\ell}}) (p_{\ell} p_{\nu_{\tau}}) ~. \end{equation} (The corresponding matrix element is also used in $\mu$ decays, but normally the $\mu$ is assumed stable.) In $\tau$ decays to hadrons, the hadrons and the $\nu_{\tau}$ are distributed according to phase space times the factor $x_{\nu} \, (3 - x_{\nu})$, where $x_{\nu} = 2E_{\nu}/m_{\tau}$ in the rest frame of the $\tau$. The latter factor is the $\nu_{\tau}$ spectrum predicted by the parton level $V-A$ matrix element, and therefore represents an attempt to take into account that the $\nu_{\tau}$ should take a larger momentum fraction than given by phase space alone. The probably largest shortcoming of the $\tau$ decay treatment is that no polarization effects are included, i.e. the $\tau$ is always assumed to decay isotropically. Usually this is not correct, since a $\tau$ is produced polarized in $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$ decays. The \ttt{LUTAUD} routine provides a generic interface to an external $\tau$ decay library, where such effects could be handled (see also \ttt{MSTJ(28)}). \subsubsection{Weak decays of bottom hadrons} Some exclusive branching ratios now start to be known for $\mathrm{B}$ decays. In this version, the $\mathrm{B}^0$, $\mathrm{B}^+$, $\mathrm{B}_{\mathrm{s}}^0$ and $\Lambda_{\b}^0$ therefore appear in a similar vein to the one outlined above for $\mathrm{D}_{\mathrm{s}}^+$ and $\Lambda_{\c}^+$ above. That is, all leptonic channels and all hadronic two-body decay channels are explicitly listed, while hadronic channels with three or more particles are only given in terms of a quark content. The $\mathrm{B}_{\c}$ is exceptional, in that either the bottom or the charm quark may decay first, and in that annihilation graphs may be non-negligible. Leptonic and semileptonic channels are here given in full, while hadronic channels are only listed in terms of a quark content, with a relative composition as given in \cite{Lus91}. No separate branching ratios are set for any of the other weakly decaying bottom hadrons, but instead a pure `spectator quark' model is assumed, where the decay of the $\b$ quark is the same in all hadrons and the only difference in final flavour content comes from the spectator quark. Compared to the charm decays, the weak decay matrix elements are given somewhat larger importance in the hadronic decay channels. In semileptonic decays $\b \to \c \ell^- \br{\nu}_{\ell}$ the $\c$ quark is combined with the spectator antiquark or diquark to form one single hadron. This hadron may be either a pseudoscalar, a vector or a higher resonance (tensor etc.). The relative fraction of the higher resonances has been picked to be about 30\%, in order to give a leptonic spectrum in reasonable experiment with data. (This only applies to the main particles $\mathrm{B}^0$, $\mathrm{B}^+$, $\mathrm{B}_{\mathrm{s}}^0$ and $\Lambda_{\b}^0$; for the rest the choice is according to the standard composition in the fragmentation.) The overall process is therefore $H \to h \ell^- \br{\nu}_{\ell}$, where $H$ is a bottom antimeson or a bottom baryon (remember that $\br{\mathrm{B}}$ is the one that contains a $\b$ quark), and the matrix element used to distribute momenta is \begin{equation} |{\cal M}|^2 = (p_H p_{\nu}) (p_{\ell} p_h) ~. \end{equation} Again decay product masses have been neglected in the matrix element, but in the branching ratios the $\tau^- \br{\nu}_{\tau}$ channel has been reduced in rate, compared with $\mathrm{e}^- \br{\nu}_{\mathrm{e}}$ and $\mu^- \br{\nu}_{\mu}$ ones, according to the expected mass effects. No CKM-suppressed decays $\b \to \u \ell^- \br{\nu}_{\ell}$ are currently included. In most multibody hadronic decays, e.g. $\b \to \c \d \overline{\mathrm{u}}$, the $\c$ quark is again combined with the spectator flavour to form one single hadron, and thereafter the hadron and the two quark momenta are distributed according to the same matrix element as above, with $\ell^- \leftrightarrow \d$ and $\br{\nu}_{\ell} \leftrightarrow \overline{\mathrm{u}}$. The invariant mass of the two quarks is calculated next. If this mass is so low that two hadrons cannot be formed from the system, the two quarks are combined into one single hadron. Else the same kind of approach as in hadronic charm decays is adopted, wherein a multiplicity is selected, a number of hadrons are formed and thereafter momenta are distributed according to phase space. The difference is that here the charm decay product is distributed according to the $V-A$ matrix element, and only the rest of the system is assumed isotropic in its rest frame, while in charm decays all hadrons are distributed isotropically. Note that the $\c$ quark and the spectator are assumed to form one colour singlet and the $\d \overline{\mathrm{u}}$ another, separate one. It is thus assumed that the original colour assignments of the basic hard process are better retained than in charm decays. However, sometimes this will not be true, and with about 20\% probability the colour assignment is flipped around so that $\c \overline{\mathrm{u}}$ forms one singlet. (In the program, this is achieved by changing the order in which decay products are given.) In particular, the decay $\b \to \c \mathrm{s} \overline{\mathrm{c}}$ is allowed to give a $\c\overline{\mathrm{c}}$ colour-singlet state part of the time, and this state may collapse to a single $\mathrm{J}/\psi$. Two-body decays of this type are explicitly listed for $\mathrm{B}^0$, $\mathrm{B}^+$, $\mathrm{B}_{\mathrm{s}}^0$ and $\Lambda_{\b}^0$; while other $\mathrm{J}/\psi$ production channels appear from the flavour content specification. The $\mathrm{B}^0$--$\br{\mathrm{B}}^0$ and $\mathrm{B}_{\mathrm{s}}^0$--$\br{\mathrm{B}}_{\mathrm{s}}^0$ systems mix before decay. This is optionally included. With a probability \begin{equation} {\cal P}_{\mrm{flip}} = \sin^2 \left( \frac{x \, \tau} { 2\, \langle \tau \rangle} \right) \end{equation} a $\mathrm{B}$ is therefore allowed to decay like a $\br{\mathrm{B}}$, and vice versa. The mixing parameters are by default $x_{\d} = 0.7$ in the $\mathrm{B}^0$--$\br{\mathrm{B}}^0$ system and $x_{\mathrm{s}} = 10$ in the $\mathrm{B}_{\mathrm{s}}^0$--$\br{\mathrm{B}}_{\mathrm{s}}^0$ one. The generic $\mathrm{B}$ meson and baryon decay properties are stored for `particle' 85. This particle contains a description of the free $\b$ quark decay, with an instruction to find the spectator flavour according to the particle code of the actual decaying hadron. Currently baryons other than $\Lambda_{\b}^0$ are treated this way. If so desired, each hadron could be given a separate decay channel list, or all $\mathrm{B}$ hadrons could be mapped to particle 85, as used to be the case.. \subsubsection{Weak decays of top and fourth generation} As already explained in section \ref{ss:decpartcont}, heavy quarks are normally assumed to decay before they fragment. Optionally, they may be allowed to fragment before they decay. In either case, the decay itself is handled as if the heavy flavour is free. For a hadron, some of the hadron energy is reserved for the spectator quark. The decay matrix element used for $\mathrm{Q} \to \mathrm{q} \overline{\mathrm{f}} \mathrm{f}$ is \begin{equation} |{\cal M}|^2 \propto \frac{ (p_{\mathrm{Q}} p_{\overline{\mathrm{f}}}) (p_{\mathrm{f}} p_{\mathrm{q}})} {\left( (p_{\mathrm{f}} + p_{\overline{\mathrm{f}}})^2 - m_{\mathrm{W}}^2 \right)^2 + m_{\mathrm{W}}^2 \Gamma_{\mathrm{W}}^2} ~. \end{equation} Here $\mathrm{Q}$ may represent the $\t$ or any of the fourth generation quarks, $\mathrm{l}$ and $\mathrm{h}$. With trivial change of notation, the lepton $\chi$ obeys the same formula. The $\mathrm{f}\overline{\mathrm{f}}$ pair are the fermions from the $\mathrm{W}$ decay, either quarks or leptons. The program takes care of the effects of the $\mathrm{W}$ propagator, whatever the mass difference $m_{\mathrm{Q}} - m_{\mathrm{q}}$, with one proviso: the selection of the $\mathrm{q}$ flavour is done according to fixed branching ratios, and does thus not take into account the relative enhancement of a CKM-suppressed $\mathrm{q}$ due to mass effects. This would play a r\^ole around thresholds, e.g., with $m_{\t} \approx m_{\mathrm{W}}$, the $\t \to \mathrm{s}$ would be enhanced compared with $\t \to \b$. On the other hand, threshold factors are included for the choice of the $\mathrm{f}\overline{\mathrm{f}}$ fermion pair from the $\mathrm{W}$ decay. For the alternative with a rapidly decaying top quark, so that no hadron is formed, one is not close to threshold. The composition of the light flavour produced in the decay is then calculated according to the respective phase space times CKM weight. By default the $\mathrm{W}$ decays with the spin information implicit in the matrix element above, but isotropic $\mathrm{W}$ decay is an option. The $\b$ quark produced in the decay $\t \to \b\mathrm{W}^+$ may be allowed to radiate. It thereby acquires an effective mass, which means that the kinematics of the decay is changed, with energy shuffled from the $\mathrm{W}$ to the $\b$. The system containing the spectator quark will often have a mass too small to allow it to fragment like a jet system. In these cases a single particle is formed from the flavour content, with a momentum vector given by the sum of the two quark momenta. Since the energy of this particle then will come out wrong, the momenta of the other jets or leptons in the decay are modified slightly to obtain total energy conservation. (Of course, for $\chi$ decay, there is no spectator and thus no treatment of this kind.) The $\mathrm{f} \overline{\mathrm{f}}$ pair from the $\mathrm{W}$ decay is allowed to shower, i.e. emit gluons and photons according to the standard final-state radiation algorithm, including matching to first-order matrix elements. The resulting jet system is fragmented with ordinary string fragmentation --- the mass is here so high that a fragmentation description is quite appropriate. Only very rarely would the $\mathrm{W}$ mass be below the threshold for the production of a pair of particles; such kinematical configurations are rejected. \subsubsection{Other decays} For onia spin 1 resonances, decay channels into a pair of leptons are explicitly given. Hadronic decays of the $\mathrm{J}/\psi$ are simulated using the flavour generation model introduced for charm. For $\Upsilon$ a fraction of the hadronic decays is into $\mathrm{q}\overline{\mathrm{q}}$ pairs, while the rest is into $\mathrm{g}\g\mathrm{g}$ or $\mathrm{g}\g\gamma$, using the matrix elements of eq.~(\ref{ee:Upsilondec}). The $\eta_c$ and $\eta_b$ are both allowed to decay into a $\mathrm{g}\g$ pair, which then subsequently fragments. In $\Upsilon$ and $\eta_b$ decays the partons are allowed to shower before fragmentation, but energies are too low for showering to have any impact. With current bounds on the top mass, one does not expect the formation of well-defined toponium states. A complete description of the resonance structure in the threshold region is beyond the scope of the program. The approach taken for the toponium states that have been defined is to let either the $\t$ or the $\overline{\mathrm{t}}$ decay weakly first, then do the fragmentation, and subsequently let the produced antitop or top hadron decay. A better description is provided by the {\tsc{Pythia}} machinery for resonance decays. Default branching ratios are given for resonances like the $\mathrm{Z}^0$, the $\mathrm{W}^{\pm}$ or the $\H^0$. When {\tsc{Pythia}} is initialized, these numbers are replaced by branching ratios evaluated from the given masses. For $\mathrm{Z}^0$ and $\mathrm{W}^{\pm}$ the branching ratios depend only marginally on the masses assumed, while effects are large e.g. for the $\H^0$. In fact, branching ratios may vary over the Breit--Wigner resonance shape, something which is also taken into account in {\tsc{Pythia}}. Therefore the default resonance treatment of {\tsc{Jetset}} is normally not so useful, and should be avoided (except, of course, the standard $\e^+\e^- \to \gamma^* / \Z^0 \to \mathrm{q}\overline{\mathrm{q}}$ description). When it is used, a channel is selected according to the given fixed branching ratios. If the decay is into a $\mathrm{q}\overline{\mathrm{q}}$ pair, the quarks are allowed to shower and subsequently the parton system is fragmented. \clearpage \section{The JETSET Program Elements} In this section we collect information on most of the routines and common block variables found in {\tsc{Jetset}}. A few parts are discussed elsewhere; this includes the $\e^+\e^-$ routines, parton showers and event-analysis routines. In this section the emphasis is on the fragmentation and decay package, and on generic utilities for things like event listing. \subsection{Definition of Initial Configuration or Variables} With the use of the conventions described for the event record, it is possible to specify any initial jet/particle configuration. This task is simplified for a number of often occuring situations by the existence of the filling routines below. It should be noted that many users do not come in direct contact with these routines, since that is taken care of by higher-level routines for specific processes, particularly \ttt{LUEEVT} and \ttt{PYEVNT}. Several calls to the routines can be combined in the specification. In case one call is enough, the complete fragmentation/decay chain may be simulated at the same time. At each call, the value of \ttt{N} is updated to the last line used for information in the call, so if several calls are used, they should be made with increasing \ttt{IP} number, or else \ttt{N} should be redefined by hand afterwards. The routine \ttt{LUJOIN} is very useful to define the colour flow in more complicated parton configurations; thereby one can bypass the not so trivial rules for how to set the \ttt{K(I,4)} and \ttt{K(I,5)} colour-flow information. As an experiment, the routine \ttt{LUGIVE} contains a facility to set various comonblock variables in a controlled and documented fashion. \drawbox{CALL LU1ENT(IP,KF,PE,THE,PHI)}\label{p:LU1ENT} \begin{entry} \itemc{Purpose:} to add one entry to the event record, i.e. either a jet or a particle. \iteme{IP :} normally line number for the jet/particle. There are two exceptions. \\ If \ttt{IP=0}, line number 1 is used and \ttt{LUEXEC} is called. \\ If \ttt{IP<0}, line \ttt{-IP} is used, with status code \ttt{K(-IP,2)=2} rather than 1; thus a jet system may be built up by filling all but the last jet of the system with \ttt{IP<0}. \iteme{KF :} jet/particle flavour code. \iteme{PE :} jet/particle energy. If \ttt{PE} is smaller than the mass, the jet/particle is taken to be at rest. \iteme{THE, PHI :} polar and azimuthal angle for the momentum vector of the jet/particle. \end{entry} \drawbox{CALL LU2ENT(IP,KF1,KF2,PECM)}\label{p:LU2ENT} \begin{entry} \itemc{Purpose:} to add two entries to the event record, i.e. either a 2-jet system or two separate particles. \iteme{IP :} normally line number for the first jet/particle, with second in line \ttt{IP+1}. There are two exceptions. \\ If \ttt{IP=0}, lines 1 and 2 are used and \ttt{LUEXEC} is called. \\ If \ttt{IP<0}, lines \ttt{-IP} and \ttt{-IP+1} are used, with status code \ttt{K(I,1)=3}, i.e. with special colour connection information, so that a parton shower can be generated by a \ttt{LUSHOW} call, followed by a \ttt{LUEXEC} call, if so desired (only relevant for jets). \iteme{KF1, KF2 :} flavour codes for the two jets/particles. \iteme{PECM :} ($=E_{\mrm{cm}}$) the total energy of the system. \itemc{Remark:} the system is given in the c.m. frame, with the first jet/particle going out in the $+z$ direction. \end{entry} \drawbox{CALL LU3ENT(IP,KF1,KF2,KF3,PECM,X1,X3)}\label{p:LU3ENT} \begin{entry} \itemc{Purpose:} to add three entries to the event record, i.e. either a 3-jet system or three separate particles. \iteme{IP :} normally line number for the first jet/particle, with other two in \ttt{IP+1} and \ttt{IP+2}. There are two exceptions. \\ If \ttt{IP=0}, lines 1, 2 and 3 are used and \ttt{LUEXEC} is called. \\ If \ttt{IP<0}, lines \ttt{-IP} through \ttt{-IP+2} are used, with status code \ttt{K(I,1)=3}, i.e. with special colour connection information, so that a parton shower can be generated by a \ttt{LUSHOW} call, followed by a \ttt{LUEXEC} call, if so desired (only relevant for jets). \iteme{KF1, KF2, KF3:} flavour codes for the three jets/particles. \iteme{PECM :} ($E_{\mrm{cm}}$) the total energy of the system. \iteme{X1, X3 :} $x_i = 2E_i/E_{\mrm{cm}}$, i.e. twice the energy fraction taken by the $i$'th jet. Thus $x_2 = 2 - x_1 - x_3$, and need not be given. Note that not all combinations of $x_i$ are inside the physically allowed region. \itemc{Remark :} the system is given in the c.m. frame, in the $xz$-plane, with the first jet going out in the $+z$ direction and the third one having $p_x > 0$. \end{entry} \drawbox{CALL LU4ENT(IP,KF1,KF2,KF3,KF4,PECM,X1,X2,X4,X12,X14)}% \label{p:LU4ENT} \begin{entry} \itemc{Purpose:} to add four entries to the event record, i.e. either a 4-jet system or four separate particles (or, for $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}'\overline{\mathrm{q}}'$ events, two 2-jet systems). \iteme{IP :} normally line number for the first jet/particle, with other three in lines \ttt{IP+1}, \ttt{IP+2} and \ttt{IP+3}. There are two exceptions. \\ If \ttt{IP=0}, lines 1, 2, 3 and 4 are used and \ttt{LUEXEC} is called. \\ If \ttt{IP<0}, lines \ttt{-IP} through \ttt{-IP+3} are used, with status code \ttt{K(I,1)=3}, i.e. with special colour connection information, so that a parton shower can be generated by a \ttt{LUSHOW} call, followed by a \ttt{LUEXEC} call, if so desired (only relevant for jets). \iteme{KF1,KF2,KF3,KF4 :} flavour codes for the four jets/particles. \iteme{PECM :} ($=E_{\mrm{cm}}$) the total energy of the system. \iteme{X1,X2,X4 :} $x_i = 2E_i/E_{\mrm{cm}}$, i.e. twice the energy fraction taken by the $i$'th jet. Thus $x_3 = 2 - x_1 - x_2 - x_4$, and need not be given. \iteme{X12,X14 :} $x_{ij} = 2 p_i p_j/E_{\mrm{cm}}^2$, i.e. twice the four-vector product of the momenta for jets $i$ and $j$, properly normalized. With the masses known, other $x_{ij}$ may be constructed from the $x_i$ and $x_{ij}$ given. Note that not all combinations of $x_i$ and $x_{ij}$ are inside the physically allowed region. \itemc{Remark:} the system is given in the c.m. frame, with the first jet going out in the $+z$ direction and the fourth jet lying in the $xz$-plane with $p_x > 0$. The second jet will have $p_y > 0$ and $p_y < 0$ with equal probability, with the third jet balancing this $p_y$ (this corresponds to a random choice between the two possible stereoisomers). \end{entry} \drawbox{CALL LUJOIN(NJOIN,IJOIN)}\label{p:LUJOIN} \begin{entry} \itemc{Purpose:} to connect a number of previously defined partons into a string configuration. Initially the partons must be given with status codes \ttt{K(I,1)=} 1, 2 or 3. Afterwards the partons all have status code 3, i.e. are given with full colour-flow information. Compared to the normal way of defining a parton system, the partons need therefore not appear in the same sequence in the event record as they are assumed to do along the string. It is also possible to call \ttt{LUSHOW} for all or some of the entries making up the string formed by \ttt{LUJOIN}. \iteme{NJOIN:} the number of entries that are to be joined by one string. \iteme{IJOIN:} an one-dimensional array, of size at least \ttt{NJOIN}. The \ttt{NJOIN} first numbers are the positions of the partons that are to be joined, given in the order the partons are assumed to appear along the string. If the system consists entirely of gluons, the string is closed by connecting back the last to the first entry. \itemc{Remarks:} only one string (i.e. one colour singlet) may be defined per call, but one is at liberty to use any number of \ttt{LUJOIN} calls for a given event. The program will check that the parton configuration specified makes sense, and not take any action unless it does. Note, however, that an initially sensible parton configuration may become nonsensical, if only some of the partons are reconnected, while the others are left unchanged. \end{entry} \drawbox{CALL LUGIVE(CHIN)}\label{p:LUGIVE} \begin{entry} \itemc{Purpose:} to set the value of any variable residing in the commmonblocks \ttt{LUJETS}, \ttt{LUDAT1}, \ttt{LUDAT2}, \ttt{LUDAT3}, \ttt{LUDAT4}, \ttt{LUDATR}, \ttt{PYSUBS}, \ttt{PYPARS}, \ttt{PYINT1}, \ttt{PYINT2}, \ttt{PYINT3}, \ttt{PYINT4}, \ttt{PYINT5}, \ttt{PYINT6}, or \ttt{PYINT7}. This is done in a more controlled fashion than by directly including the common blocks in the user program, in that array bounds are checked and the old and new values for the variable changed are written to the output for reference. \iteme{CHIN :} character expression of length at most 100 characters, with requests for variables to be changed, stored in the form \\ \ttt{variable1=value1;variable2=value2;variable3=value3}\ldots~. \\ Note that an arbitrary number of instructions can be stored in one call if separated by semicolons, and that blanks may be included anyplace. The variable$_i$ may be any single variable in the {\tsc{Jetset/Pythia}} common blocks, and the value$_i$ must be of the correct integer, real or character (without extra quotes) type. Array indices and values must be given explicitly, i.e. cannot be variables in their own right. The exception is that the first index can be preceded by a \ttt{C}, signifying that the index should be translated from normal KF to compressed KC code with a \ttt{LUCOMP} call; this is allowed for the \ttt{KCHG}, \ttt{PMAS}, \ttt{MDCY} and \ttt{CHAF} arrays. If a value$_i$ is omitted, i.e. with the construction \ttt{variable=}, the current value is written to the output, but the variable itself is not changed. \itemc{Remark :} The checks on array bounds are hardwired into this routine. Therefore, if some user changes array dimensions and \ttt{MSTU(3)}, \ttt{MSTU(6)} and/or \ttt{MSTU(7)}, as allowed by other considerations, these changes will not be known to \ttt{LUGIVE}. Normally this should not be a problem, however. \end{entry} \subsection{The JETSET Physics Routines} \label{ss:JETphysrout} The physics routines form the major part of {\tsc{Jetset}}, but once the initial jet/particle configuration has been specified and default parameter values changed, if so desired, only a \ttt{LUEXEC} call is necessary to simulate the whole fragmentation and decay chain. Therefore a normal user will not directly see any of the other routines in this section. Some of them could be called directly, but the danger of faulty usage is then non-negligible. The \ttt{LUTAUD} routine provides an optional interface to an external $\tau$ decay library, where polarization effects could be included. It is up to the user to write the appropriate calls, as explained at the end of this section. \drawbox{CALL LUEXEC}\label{p:LUEXEC} \begin{entry} \itemc{Purpose:} to administrate the fragmentation and decay chain. \ttt{LUEXEC} may be called several times, but only entries which have not yet been treated (more precisely, which have $1 \leq$\ttt{K(I,1)}$\leq 10$) can be affected by further calls. This may apply if more jets/particles have been added by the user, or if particles previously considered stable are now allowed to decay. The actions that will be taken during a \ttt{LUEXEC} call can be tailored extensively via the \ttt{LUDAT1}--\ttt{LUDAT3} common blocks, in particular by setting the \ttt{MSTJ} values suitably. \end{entry} \boxsep \begin{entry} \iteme{SUBROUTINE LUPREP(IP) :}\label{p:LUPREP} to rearrange parton shower end products (marked with \ttt{K(I,1)=3}) sequentially along strings; also to (optionally) allow small jet systems to collapse into two particles or one only, in the latter case with energy and momentum to be shuffled elsewhere in the event; also to perform checks that e.g. flavours of colour-singlet systems make sense. \iteme{SUBROUTINE LUSTRF(IP) :}\label{p:LUSTRF} to generate the fragmentation of an arbitrary colour-singlet jet system according to the Lund string fragmentation model. In many respects, this routine is the very heart and soul of {\tsc{Jetset}}. \iteme{SUBROUTINE LUINDF(IP) :}\label{p:LUINDF} to handle the fragmentation of a jet system according to independent fragmentation models, and implement energy, momentum and flavour conservation, if so desired. Also the fragmentation of a single jet, not belonging to a jet system, is considered here (this is of course physical nonsense, but may sometimes be convenient for specific tasks). \iteme{SUBROUTINE LUDECY(IP) :}\label{p:LUDECY} to perform a particle decay, according to known branching ratios or different kinds of models, depending on our level of knowledge. Various matrix elements are included for specific processes. \iteme{SUBROUTINE LUKFDI(KFL1,KFL2,KFL3,KF) :}\label{p:LUKFDI} to generate a new quark or diquark flavour and to combine it with an existing flavour to give a hadron. \begin{subentry} \iteme{KFL1:} incoming flavour. \iteme{KFL2:} extra incoming flavour, e.g. for formation of final particle, where the flavours are completely specified. Is normally 0. \iteme{KFL3:} newly created flavour; is 0 if \ttt{KFL2} is non-zero. \iteme{KF:} produced hadron. Is 0 if something went wrong (e.g. inconsistent combination of incoming flavours). \end{subentry} \iteme{SUBROUTINE LUPTDI(KFL,PX,PY) :}\label{p:LUPTDI} to give transverse momentum, e.g. for a $\mathrm{q}\overline{\mathrm{q}}$ pair created in the colour field, according to independent Gaussian distributions in $p_x$ and $p_y$. \iteme{SUBROUTINE LUZDIS(KFL1,KFL3,PR,Z) :}\label{p:LUZDIS} to generate the longitudinal scaling variable $z$ in jet fragmentation, either according to the Lund symmetric fragmentation function, or according to a choice of other shapes. \iteme{SUBROUTINE LUBOEI :}\label{p:LUBOEI} to include Bose--Einstein effects according to a simple para\-metri\-zation. By default, this routine is not called. If called from \ttt{LUEXEC}, this is done after the decay of short-lived resonances, but before the decay of long-lived ones. This means the routine should never be called directly by you, nor would effects be correctly simulated if decays are switched off. See \ttt{MSTJ(51) - MSTJ(52)} for switching on the routine. \iteme{FUNCTION ULMASS(KF) :}\label{p:ULMASS} to give the mass for a parton/particle. \iteme{SUBROUTINE LUNAME(KF,CHAU) :}\label{p:LUNAME} to give the parton/particle name (as a string of type \ttt{CHARACTER CHAU*16}). \iteme{FUNCTION LUCHGE(KF) :}\label{p:LUCHGE} to give three times the charge for a parton/particle. \iteme{FUNCTION LUCOMP(KF) :}\label{p:LUCOMP} to give the compressed parton/particle code KC for a given KF code, as required to find entry into mass and decay data tables. Also checks whether the given KF code is actually an allowed one (i.e. known by the program), and returns 0 if not. Note that KF may be positive or negative, while the resulting KC code is never negative. \iteme{SUBROUTINE LUERRM(MERR,MESSAG) :}\label{p:LUERRM} to keep track of the number of errors and warnings encountered, write out information on them, and abort the program in case of too many errors. \iteme{FUNCTION ULANGL(X,Y) :}\label{p:ULANGL} to calculate the angle from the $x$ and $y$ coordinates. \iteme{SUBROUTINE LULOGO :}\label{p:LULOGO} to write a titlepage for the {\tsc{Jetset/Pythia}} programs. Called by LULIST(0). \iteme{BLOCK DATA LUDATA :}\label{p:LUDATA} to give default values for variables in the \ttt{LUDAT1}, \ttt{LUDAT2}, \ttt{LUDAT3}, \ttt{LUDAT4} and \ttt{LUDATR} common blocks. \end{entry} \drawbox{CALL LUTAUD(ITAU,IORIG,KFORIG,NDECAY)}\label{p:LUTAUD} \begin{entry} \itemc{Purpose:} to act as an interface between the standard decay routine \ttt{LUDECY} and a user-supplied $\tau$ lepton decay library. The latter library would normally know how to handle polarized $\tau$'s, given the $\tau$ polarization as input, so one task of the interface routine is to construct the $\tau$ polarization/helicity from the information available. Input to the routine (from \ttt{LUDECY}) is provided in the first three arguments, while the last argument and some event record information have to be set before return. To use this facility you have to set the switch \ttt{MSTJ(28)}, include your own interface routine \ttt{LUTAUD} and see to it that the dummy routine \ttt{LUTAUD} in {\tsc{Jetset}} is not linked. The dummy routine is there only to avoid unresolved external references when no user-supplied interface is linked. \iteme{ITAU :} line number in the event record where the $\tau$ is stored. The four-momentum of this $\tau$ has first been boosted back to the rest frame of the decaying mother and thereafter rotated to move out along the $+z$ axis. It would have been possible to also perform a final boost to the rest frame of the $\tau$ itself, but this has been avoided so as not to suppress the kinematics aspect of close-to-threshold production (e.g. in $\mathrm{B}$ decays) vs. high-energy production (e.g. in real $\mathrm{W}$ decays). The choice of frame should help the calculation of the helicity configuration. After the \ttt{LUTAUD} call the $\tau$ and its decay products will automatically be rotated and boosted back. However, seemingly, the event record does not conserve momentum at this intermediate stage. \iteme{IORIG :} line number where the mother particle to the $\tau$ is stored. Is 0 if the mother is not stored. This does not have to mean the mother is unknown. For instance, in semileptonic $\mathrm{B}$ decays the mother is a $\mathrm{W}^{\pm}$ with known four-momentum $p_{\mathrm{W}} = p_{\tau} + p_{\nu_{\tau}}$, but there is no $\mathrm{W}$ line in the event record. When several copies of the mother is stored (e.g. one in the documentation section of the event record and one in the main section), \ttt{IORIG} points to the last. If a branchings like $\tau \to \tau\gamma$ occurs, the `grandmother' is given, i.e. the mother of the direct $\tau$ before branching. \iteme{KFORIG :} flavour code for the mother particle. Is 0 if the mother is unknown. The mother would typically be a resonance such as $\gamma^* / \Z^0$ (23), $\mathrm{W}^{\pm}$ ($\pm 24$), $\H^0$ (25), or $\H^{\pm}$ ($\pm 37$). Often the helicity choice would be clear just by the knowledge of this mother species, e.g., $\mathrm{W}^{\pm}$ vs. $\H^{\pm}$. However, sometimes further complications may exist. For instance, the KF code 23 represents a mixture of $\gamma^*$ and $\mathrm{Z}^0$; a knowledge of the mother mass (in \ttt{P(IORIG,5)}) would here be required to make the choice of helicities. Further, a $\mathrm{W}^{\pm}$ or $\mathrm{Z}^0$ may either be (predominantly) transverse or longitudinal, depending on the production process under study. \iteme{NDECAY :} the number of decay products of the $\tau$; to be given by the user. You must also store the KF flavour codes of those decay products in the positions \ttt{K(I,2)}, \ttt{N+1}$\leq$\ttt{I}$\leq$\ttt{N+NDECAY}, of the event record. The corresponding five-momentum (momentum, energy and mass) should be stored in the associated \ttt{P(I,J)} positions, 1$\leq$\ttt{J}$\leq$5. The four-momenta are expected to add up to the four-momentum of the $\tau$ in position \ttt{ITAU}. You should not change the \ttt{N} value or any of the other \ttt{K} or \ttt{V} values (neither for the $\tau$ nor for its decay products) since this is automatically done in \ttt{LUDECY}. \end{entry} \subsection{Event Study and Data Listing Routines} After an \ttt{LUEXEC} call, the event generated is stored in the \ttt{LUJETS} common block, and whatever physical variable is desired may be constructed from this record. An event may be rotated, boosted or listed, and particle data may be listed or modified. Via the functions \ttt{KLU} and \ttt{PLU} the values of some frequently appearing variables may be obtained more easily. As described in section \ref{ss:evanrout}, also more detailed event shape analyses may be performed simply. \drawbox{CALL LUROBO(THE,PHI,BEX,BEY,BEZ)}\label{p:LUROBO} \begin{entry} \itemc{Purpose:} to perform rotations and Lorentz boosts (in that order, if both in the same call) of jet/particle momenta and vertex position variables. \iteme{THE, PHI :} standard polar coordinates $\theta, \varphi$, giving the rotated direction of a momentum vector initially along the $+z$ axis. \iteme{BEX, BEY, BEZ :} gives the direction and size \mbox{\boldmath $\beta$} of a Lorentz boost, such that a particle initially at rest will have $\mbf{p}/E = $\mbox{\boldmath $\beta$} afterwards. \itemc{Remark:} all entries 1 through \ttt{N} are affected by the transformation, unless lower and upper bounds are explicitly given by \ttt{MSTU(1)} and \ttt{MSTU(2)}, or if status code \ttt{K(I,1)}$\leq 0$. \end{entry} \drawbox{ENTRY LUDBRB(IMI,IMA,THE,PHI,DBEX,DBEY,DBEZ)}\label{p:LUDBRB} \begin{entry} \itemc{Purpose:} to perform rotations and Lorentz boosts (in that order, if both in the same call) of jet/particle momenta and vertex position variables, for a specific range of entries, and with the boost vector given in double precision. Is entry to \ttt{LUROBO}, mainly intended for internal use. \iteme{IMI, IMA :} range of entries affected by transformation, \ttt{IMI}$\leq$\ttt{I}$\leq$\ttt{IMA}. \iteme{THE, PHI :} standard polar coordinates $\theta, \varphi$, giving the rotated direction of a momentum vector initially along the $+z$ axis. \iteme{DBEX, DBEY, DBEZ :} gives the direction and size \mbox{\boldmath $\beta$} of a Lorentz boost, such that a particle initially at rest will have $\mbf{p}/E = $\mbox{\boldmath $\beta$} afterwards. Is to be given in double precision. \itemc{Remark:} all entries with status codes \ttt{K(I,1)>0} in the requested range are affected by the transformation. \end{entry} \drawbox{CALL LUEDIT(MEDIT)}\label{p:LUEDIT} \begin{entry} \itemc{Purpose:} to exclude unstable or undetectable jets/particles from the event record. One may also use \ttt{LUEDIT} to store spare copies of events (specifically initial parton configuration) that can be recalled to allow e.g. different fragmentation schemes to be run through with one and the same parton configuration. Finally, an event which has been analyzed with \ttt{LUSPHE}, \ttt{LUTHRU} or \ttt{LUCLUS} (see section \ref{ss:evanrout}) may be rotated to align the event axis with the $z$ direction. \iteme{MEDIT :} tells which action is to be taken. \begin{subentry} \iteme{= 0 :} empty (\ttt{K(I,1)=0}) and documentation (\ttt{K(I,1)>20}) lines are removed. The jets/particles remaining are compressed in the beginning of the \ttt{LUJETS} common block and the \ttt{N} value is updated accordingly. The event history is lost, so that information stored in \ttt{K(I,3)}, \ttt{K(I,4)} and \ttt{K(I,5)} is no longer relevant. \iteme{= 1 :} as \ttt{=0}, but in addition all jets/particles that have fragmented/decayed (\ttt{K(I,1)>10}) are removed. \iteme{= 2 :} as \ttt{=1}, but also all neutrinos and unknown particles (i.e. compressed code KC$ = 0$) are removed. \iteme{= 3 :} as \ttt{=2}, but also all uncharged, colour neutral particles are removed, leaving only charged, stable particles (and unfragmented partons, if fragmentation has not been performed). \iteme{= 5 :} as \ttt{=0}, but also all partons which have branched or been rearranged in a parton shower and all particles which have decayed are removed, leaving only the fragmenting parton configuration and the final-state particles. \iteme{= 11 :} remove lines with \ttt{K(I,1)<0}. Update event history information (in \ttt{K(I,3) - K(I,5)}) to refer to remaining entries. \iteme{= 12 :} remove lines with \ttt{K(I,1)=0}. Update event history information (in \ttt{K(I,3) - K(I,5)}) to refer to remaining entries. \iteme{= 13 :} remove lines with \ttt{K(I,1)}$ = 11$, 12 or 15, except for any line with \ttt{K(I,2)=94}. Update event history information (in \ttt{K(I,3) - K(I,5)}) to refer to remaining entries. In particular, try to trace origin of daughters, for which the mother is decayed, back to entries not deleted. \iteme{= 14 :} remove lines with \ttt{K(I,1)}$ = 13$ or 14, and also any line with \ttt{K(I,2)=94}. Update event history information (in \ttt{K(I,3) - K(I,5)}) to refer to remaining entries. In particular, try to trace origin of rearranged jets back through the parton-shower history to the shower initiator. \iteme{= 15 :} remove lines with \ttt{K(I,1)>20}. Update event history information (in \ttt{K(I,3) - K(I,5)}) to refer to remaining entries. \iteme{= 16 :} try to reconstruct missing daughter pointers of decayed particles from the mother pointers of decay products. These missing pointers typically come from the need to use \ttt{K(I,4)} and \ttt{K(I,5)} also for colour flow information. \iteme{= 21 :} all partons/particles in current event record are stored (as a spare copy) in bottom of common block \ttt{LUJETS} (is e.g. done to save original partons before calling \ttt{LUEXEC}). \iteme{= 22 :} partons/particles stored in bottom of event record with \ttt{=21} are placed in beginning of record again, overwriting previous information there (so that e.g. a different fragmentation scheme can be used on the same partons). Since the copy at bottom is unaffected, repeated calls with \ttt{=22} can be made. \iteme{= 23 :} primary partons/particles in the beginning of event record are marked as not fragmented or decayed, and number of entries \ttt{N} is updated accordingly. Is simpe substitute for \ttt{=21} plus \ttt{=22} when no fragmentation/decay products precede any of the original partons/particles. \iteme{= 31 :} rotate largest axis, determined by \ttt{LUSPHE}, \ttt{LUTHRU} or \ttt{LUCLUS}, to sit along the $z$ direction, and the second largest axis into the $xz$ plane. For \ttt{LUCLUS} it can be further specified to $+z$ axis and $xz$ plane with $x > 0$, respectively. Requires that one of these routines has been called before. \iteme{= 32 :} mainly intended for \ttt{LUSPHE} and \ttt{LUTHRU}, this gives a further alignment of the event, in addition to the one implied by \ttt{=31}. The `slim' jet, defined as the side ($z > 0$ or $z < 0$) with the smallest summed $p_{\perp}$ over square root of number of particles, is rotated into the $+z$ hemisphere. In the opposite hemisphere (now $z < 0$), the side of $x > 0$ and $x < 0$ which has the largest summed $|p_z|$ is rotated into the $z < 0, x > 0$ quadrant. Requires that \ttt{LUSPHE} or \ttt{LUTHRU} has been called before. \end{subentry} \itemc{Remark:} all entries 1 through \ttt{N} are affected by the editing. For options 0--5 lower and upper bounds can be explicitly given by \ttt{MSTU(1)} and \ttt{MSTU(2)}. \end{entry} \drawbox{CALL LULIST(MLIST)}\label{p:LULIST} \begin{entry} \itemc{Purpose:} to list an event, jet or particle data, or current parameter values. \iteme{MLIST :} determines what is to be listed. \begin{subentry} \iteme{= 0 :} writes a title page, common for {\tsc{Jetset}} and {\tsc{Pythia}}, with program version numbers and last dates of change; is mostly for internal use. \iteme{= 1 :} gives a simple list of current event record, in an 80 column format suitable for viewing directly on the computer terminal. For each entry, the following information is given: the entry number \ttt{I}, the parton/particle name (see below), the status code (\ttt{K(I,1)}), the flavour code KF (\ttt{K(I,2)}), the line number of the mother (\ttt{K(I,3)}), and the three-momentum, energy and mass (\ttt{P(I,1) - P(I,5)}). If \ttt{MSTU(3)} is non-zero, lines immediately after the event record proper are also listed. A final line contains information on total charge, momentum, energy and invariant mass. \\ The particle name is given by a call to the routine \ttt{LUNAME}. For an entry which has decayed/fragmented (\ttt{K(I,1)=} 11--20), this particle name is given within parentheses. Similarly, a documentation line (\ttt{K(I,1)=} 21--30) has the name enclosed in expression signs (!\ldots!) and an event/jet axis information line the name within inequality signs ($<$\ldots$>$). If the last character of the name is a `?', it is a signal that the complete name has been truncated to fit in, and can therefore not be trusted; this is very rare. For partons which have been arranged along strings (\ttt{K(I,1)=} 1, 2, 11 or 12), the end of the parton name column contains information about the colour string arrangement: an \ttt{A} for the first entry of a string, an \ttt{I} for all intermediate ones, and a \ttt{V} for the final one (a poor man's rendering of a vertical doublesided arrow, $\updownarrow$). \\ It is possible to insert lines just consisting of sequences of \ttt{======} to separate different sections of the event record, see \ttt{MSTU(70) - MSTU(80)}. \iteme{= 2 :} gives a more extensive list of the current event record, in a 132 column format, suitable for printers or workstations. For each entry, the following information is given: the entry number \ttt{I}, the parton/particle name (with padding as described for \ttt{=1}), the status code (\ttt{K(I,1)}), the flavour code KF (\ttt{K(I,2)}), the line number of the mother (\ttt{K(I,3)}), the decay product/colour-flow pointers (\ttt{K(I,4), K(I,5)}), and the three-momentum, energy and mass (\ttt{P(I,1) - P(I,5)}). If \ttt{MSTU(3)} is non-zero, lines immediately after the event record proper are also listed. A final line contains information on total charge, momentum, energy and invariant mass. Lines with only \ttt{======} may be inserted as for \ttt{=1}. \iteme{= 3 :} gives the same basic listing as \ttt{=2}, but with an additional line for each entry containing information on production vertex position and time (\ttt{V(I,1) - V(I,4)}) and, for unstable particles, proper lifetime (\ttt{V(I,5)}). \iteme{= 11 :} provides a simple list of all parton/particle codes defined in the program, with KF code and corresponding particle name. The list is grouped by particle kind, and only within each group in ascending order. \iteme{= 12 :} provides a list of all parton/particle and decay data used in the program. Each parton/particle code is represented by one line containing KF flavour code, KC compressed code, particle name, antiparticle name (where appropriate), electrical and colour charge (stored in \ttt{KCHG}), mass, resonance width and maximum broadening, average proper lifetime (in \ttt{PMAS}) and whether the particle is considered stable or not (in \ttt{MDCY}). Immediately after a particle, each decay channel gets one line, containing decay channel number (\ttt{IDC} read from \ttt{MDCY}), on/off switch for the channel, matrix element type (\ttt{MDME}), branching ratio (\ttt{BRAT}), and decay products (\ttt{KFDP}). The \ttt{MSTU(14)} flag can be used to set the maximum flavour for which particles are listed, with the default (= 0) corresponding to separately defined ones (KC$ > 100$ if KF$ > 0$). In order to keep the size down, decay modes of heavy hadrons collectively defined are never listed; these have KC codes 84--88, where the relevant information may be found. \iteme{= 13 :} gives a list of current parameter values for \ttt{MSTU}, \ttt{PARU}, \ttt{MSTJ} and \ttt{PARJ}, and the first 200 entries of \ttt{PARF}. This is useful to keep check of which default values were changed in a given run. \end{subentry} \itemc{Remark:} for options 1--3 and 12 lower and upper bounds of the listing can be explicitly given by \ttt{MSTU(1)} and \ttt{MSTU(2)}. \end{entry} \drawbox{CALL LUUPDA(MUPDA,LFN)}\label{p:LUUPDA} \begin{entry} \itemc{Purpose:} to give you the ability to update particle data, or to keep several versions of modified particle data for special purposes (e.g. charm studies). \iteme{MUPDA :} gives the type of action to be taken. \begin{subentry} \iteme{= 1 :} write a table of particle data, that you then can edit at leisure. For ordinary listing of decay data, \ttt{LULIST(12)} should be used, but that listing could not be read back in by the program. \\ For each compressed flavour code KC = 1--500, one line is written containing KC (\ttt{I5}), the basic particle name (i.e. excluding charge etc.) (\ttt{2X,A8}) in \ttt{CHAF}, the electric (\ttt{I3}), colour charge (\ttt{I3}) and particle/antiparticle distinction (\ttt{I3}) codes in \ttt{KCHG}, the mass (\ttt{F12.5}), the mass width (\ttt{F12.5}), maximum broadening (\ttt{F12.5}) and average proper lifetime (\ttt{2X,F12.5}) in \ttt{PMAS}, and the on/off decay switch (I3) in \ttt{MDCY(KC,1).} \\ After a KC line follows one line for each possible decay channel, containing the \ttt{MDME} codes (\ttt{5X,2I5}), the branching ratio (\ttt{5X,F12.5}) in \ttt{BRAT}, and the \ttt{KFDP} codes for the decay products (\ttt{5I8}), with trailing 0's if the number of decay products is smaller than 5. \iteme{= 2 :} read in particle data, as written with \ttt{=1} and thereafter edited by you, and use this data subsequently in the current run. Reading is done with fixed format, which means that you have to preserve the format codes described for \ttt{=1} during the editing. A number of checks will be made to see if input looks reasonable, with warnings if not. If some decay channel is said not to conserve charge, it should be taken seriously. Warnings that decay is kinematically unallowed need not be as serious, since that particular decay mode may not be switched on unless the particle mass is increased. \iteme{= 3 :} write current particle data as data lines, which can be edited into \ttt{BLOCK DATA LUDATA} for a permanent replacement of the particle data. This option is intended for the program author only, not for you. \end{subentry} \iteme{LFN :} the file number which the data should be written to or read from. You must see to it that this file is properly opened for read or write (since the definition of file names is machine dependent). \end{entry} \drawbox{KK = KLU(I,J)}\label{p:KLU} \begin{entry} \itemc{Purpose:} to provide various integer-valued event data. Note that many of the options available (in particular \ttt{I}$ > 0$, \ttt{J}$\geq 14$) which refer to event history will not work after a \ttt{LUEDIT} call. Further, the options 14--18 depend on the way the event history has been set up, so with the explosion of different allowed formats these options are no longer as safe as they may have been. For instance, option 16 can only work if \ttt{MSTU(16)=2}. \iteme{I=0, J= :} properties referring to the complete event. \begin{subentry} \iteme{= 1 :} \ttt{N}, total number of lines in event record. \iteme{= 2 :} total number of partons/particles remaining after fragmentation and decay. \iteme{= 6 :} three times the total charge of remaining (stable) partons and particles. \end{subentry} \iteme{I>0, J= :} properties referring to the entry in line no. \ttt{I} of the event record. \begin{subentry} \iteme{= 1 - 5 :} \ttt{K(I,1) - K(I,5)}, i.e. parton/particle status KS, flavour code KF and origin/decay product/colour-flow information. \iteme{= 6 :} three times parton/particle charge. \iteme{= 7 :} 1 for a remaining entry, 0 for a decayed, fragmented or documentation entry. \iteme{= 8 :} KF code (\ttt{K(I,2)}) for a remaining entry, 0 for a decayed, fragmented or documentation entry. \iteme{= 9 :} KF code (\ttt{K(I,2)}) for a parton (i.e. not colour neutral entry), 0 for a particle. \iteme{= 10 :} KF code (\ttt{K(I,2)}) for a particle (i.e. colour neutral entry), 0 for a parton. \iteme{= 11 :} compressed flavour code KC. \iteme{= 12 :} colour information code, i.e. 0 for colour neutral, 1 for colour triplet, -1 for antitriplet and 2 for octet. \iteme{= 13 :} flavour of `heaviest' quark or antiquark (i.e. with largest code) in hadron or diquark (including sign for antiquark), 0 else. \iteme{= 14 :} generation number. Beam particles or virtual exchange particles are generation 0, original jets/particles generation 1 and then 1 is added for each step in the fragmentation/decay chain. \iteme{= 15 :} line number of ancestor, i.e. predecessor in first generation (generation 0 entries are disregarded). \iteme{= 16 :} rank of a hadron in the jet it belongs to. Rank denotes the ordering in flavour space, with hadrons containing the original flavour of the jet having rank 1, increasing by 1 for each step away in flavour ordering. All decay products inherit the rank of their parent. Whereas the meaning of a first-rank hadron in a quark jet is always well-defined, the definition of higher ranks is only meaningful for independently fragmenting quark jets. In other cases, rank refers to the ordering in the actual simulation, which may be of little interest. \iteme{= 17 :} generation number after a collapse of a jet system into one particle, with 0 for an entry not coming from a collapse, and -1 for entry with unknown history. A particle formed in a collapse is generation 1, and then one is added in each decay step. \iteme{= 18 :} number of decay/fragmentation products (only defined in a collective sense for fragmentation). \iteme{= 19 :} origin of colour for showering parton, 0 else. \iteme{= 20 :} origin of anticolour for showering parton, 0 else. \iteme{= 21 :} position of colour daughter for showering parton, 0 else. \iteme{= 22 :} position of anticolour daughter for showering parton, 0 else. \end{subentry} \end{entry} \drawbox{PP = PLU(I,J)}\label{p:PLU} \begin{entry} \itemc{Purpose:} to provide various real-valued event data. Note that some of the options available (\ttt{I}$ > 0$, \ttt{J}$ = 20$--25), which are primarily intended for studies of systems in their respective c.m. frame, requires that a \ttt{LUEXEC} call has been made for the current initial parton/particle configuration, but that the latest \ttt{LUEXEC} call has not been followed by a \ttt{LUROBO} one. \iteme{I=0, J= :} properties referring to the complete event. \begin{subentry} \iteme{= 1 - 4 :} sum of $p_x$, $p_y$, $p_z$ and $E$, respectively, for the stable remaining entries. \iteme{= 5 :} invariant mass of the stable remaining entries. \iteme{= 6 :} sum of electric charge of the stable remaining entries. \end{subentry} \iteme{I>0, J= :} properties referring to the entry in line no. \ttt{I} of the event record. \begin{subentry} \iteme{= 1 - 5 :} \ttt{P(I,1) - P(I,5)}, i.e. normally $p_x$, $p_y$, $p_z$, $E$ and $m$ for jet/particle. \iteme{= 6 :} electric charge $e$. \iteme{= 7 :} squared momentum $|\mbf{p}|^2 = p_x^2 + p_y^2 + p_z^2$. \iteme{= 8 :} absolute momentum $|\mbf{p}|$. \iteme{= 9 :} squared transverse momentum $p_{\perp}^2 = p_x^2 + p_y^2$. \iteme{= 10 :} transverse momentum $p_{\perp}$. \iteme{= 11 :} squared transverse mass $m_{\perp}^2 = m^2 + p_x^2 + p_y^2$. \iteme{= 12 :} transverse mass $m_{\perp}$. \iteme{= 13 - 14 :} polar angle $\theta$ in radians (between 0 and $\pi$) or degrees, respectively. \iteme{= 15 - 16 :} azimuthal angle $\varphi$ in radians (between $-\pi$ and $\pi$) or degrees, respectively. \iteme{= 17 :} true rapidity $y = (1/2) \, \ln((E+p_z)/(E-p_z))$. \iteme{= 18 :} rapidity $y_{\pi}$ obtained by assuming that the particle is a pion when calculating the energy $E$, to be used in the formula above, from the (assumed known) momentum $\mbf{p}$. \iteme{= 19 :} pseudorapidity $\eta = (1/2) \, \ln((p+p_z)/(p-p_z))$. \iteme{= 20 :} momentum fraction $x_p = 2|\mbf{p}|/W$, where $W$ is the total energy of initial jet/particle configuration. \iteme{= 21 :} $x_{\mrm{F}} = 2p_z/W$ (Feynman-$x$ if system is studied in the c.m. frame). \iteme{= 22 :} $x_{\perp} = 2p_{\perp}/W$. \iteme{= 23 :} $x_E = 2E/W$. \iteme{= 24 :} $z_+ = (E+p_z)/W$. \iteme{= 25 :} $z_- = (E-p_z)/W$. \end{subentry} \end{entry} \subsection{The General Switches and Parameters} \label{ss:JETswitch} The common block \ttt{LUDAT1} is, next to \ttt{LUJETS}, the one a {\tsc{Jetset}} user is most likely to access. Here one may control in detail what the program is to do, if the default mode of operation is not satisfactory. \drawbox{COMMON/LUDAT1/MSTU(200),PARU(200),MSTJ(200),PARJ(200)}% \label{p:LUDAT1} \begin{entry} \itemc{Purpose:} to give access to a number of status codes and parameters which regulate the performance of the program as a whole. Here \ttt{MSTU} and \ttt{PARU} are related to utility functions, as well as a few parameters of the Standard Model, while \ttt{MSTJ} and \ttt{PARJ} affect the underlying physics assumptions. Some of the variables in \ttt{LUDAT1} are described elsewhere, and are therefore here only reproduced as references to the relevant sections. This in particular applies to many coupling constants mainly used by {\tsc{Pythia}}, which are found just after this, in section \ref{ss:coupcons}. \boxsep \iteme{MSTU(1),MSTU(2) :}\label{p:MSTU} (D=0,0) can be used to replace the ordinary lower and upper limits (normally 1 and \ttt{N}) for the action of \ttt{LUROBO}, and most \ttt{LUEDIT} and \ttt{LULIST} calls. Are reset to 0 in a \ttt{LUEXEC} call. \iteme{MSTU(3) :} (D=0) number of lines with extra information added after line \ttt{N}. Is reset to 0 in a \ttt{LUEXEC} call, or in an \ttt{LUEDIT} call when particles are removed. \iteme{MSTU(4) :} (D=4000) number of lines available in the common block \ttt{LUJETS}. Should always be changed if the dimensions of the \ttt{K} and \ttt{P} arrays are changed by the user, but should otherwise never be touched. Maximum allowed value is 10000, unless \ttt{MSTU(5)} is also changed. \iteme{MSTU(5) :} (D=10000) is used in building up the special colour-flow information stored in \ttt{K(I,4)} and \ttt{K(I,5)} for \ttt{K(I,3)=} 3, 13 or 14. The generic form for \ttt{j=} 4 or 5 is \\ \ttt{K(I,j)}$ = 2 \times$\ttt{MSTU(5)}$^2 \times$MCFR$ + $\ttt{MSTU(5)}$^2 \times$MCTO$ + $\ttt{MSTU(5)}$\times$ICFR$ + $ICTO, \\ with notation as in section \ref{ss:evrec}. One should always have \ttt{MSTU(5)}$\geq$\ttt{MSTU(4)}. On a 32 bit machine, values \ttt{MSTU(5)}$> 20000$ may lead to overflow problems, and should be avoided. \iteme{MSTU(6) :} (D=500) number of KC codes available in the \ttt{KCHG}, \ttt{PMAS}, \ttt{MDCY}, and \ttt{CHAF} arrays; should be changed if these dimensions are changed. \iteme{MSTU(7) :} (D=2000) number of decay channels available in the \ttt{MDME}, \ttt{BRAT} and \ttt{KFDP} arrays; should be changed if these dimensions are changed. \iteme{MSTU(10) :} (D=2) use of parton/particle masses in filling routines (\ttt{LU1ENT}, \ttt{LU2ENT}, \ttt{LU3ENT}, \ttt{LU4ENT}). \begin{subentry} \iteme{= 0 :} assume the mass to be zero. \iteme{= 1 :} keep the mass value stored in \ttt{P(I,5)}, whatever it is. (This may be used e.g. to describe kinematics with off-mass-shell partons). \iteme{= 2 :} find masses according to mass tables as usual. \end{subentry} \iteme{MSTU(11) :} (D=6) file number to which all program output is directed. It is your responsibility to see to it that the corresponding file is also opened for output. \iteme{MSTU(12) :} (D=1) writing of title page (version number and last date of change for {\tsc{Pythia}} and {\tsc{Jetset}}) on output file. \begin{subentry} \iteme{= 0 :} not done. \iteme{= 1 :} title page is written at first occasion, at which time \ttt{MSTU(12)} is set =0. \end{subentry} \iteme{MSTU(13) :} (D=1) writing of information on variable values changed by a \ttt{LUGIVE} call. \begin{subentry} \iteme{= 0 :} no information is provided. \iteme{= 1 :} information is written to standard output. \end{subentry} \iteme{MSTU(14) :} (D=0) if non-zero, this gives the maximum flavour for which a \ttt{LULIST(12)} call will give particle data on possible hadrons. With \ttt{MSTU(14)=5} only known hadrons, i.e. up to bottom, are listed. If \ttt{=0}, only separately specified particles are listed (i.e. either KF$\leq 100$ or else both KF$> 100$ and KC$> 100$). \iteme{MSTU(15) :} (D=1) selection for characters used in particle names to denote an antiparticle; appear in \ttt{LULIST} listings or other \ttt{LUNAME} applications. \begin{subentry} \iteme{= 1 :} the tilde character `$\sim$'. \iteme{= 2 :} the characters `bar'. \end{subentry} \iteme{MSTU(16) :} (D=1) choice of mother pointers for the particles produced by a fragmenting parton system. \begin{subentry} \iteme{= 1 :} all primary particles of a system point to a line with KF = 92 or 93, for string or independent fragmentation, respectively, or to a line with KF = 91 if a jet system has so small a mass that it is forced to decay into one or two particles. The two (or more) shower initiators of a showering parton system point to a line with KF = 94. The entries with KF = 91--94 in their turn point back to the predecessor partons, so that the KF = 91--94 entries form a part of the event history proper. \iteme{= 2 :} although the lines with KF = 91--94 are present, and contain the correct mother and daughter pointers, they are not part of the event history proper, in that particles produced in string fragmentation point directly to either of the two endpoint partons of the string (depending on the side they were generated from), particles produced in independent fragmentation point to the respective parton they were generated from, particles in small mass systems point to either endpoint parton, and shower initiators point to the original on-mass-shell counterparts. Also the daugher pointers bypass the KF = 91--94 entries. In independent fragmentation, a parton need not produce any particles at all, and then have daughter pointers 0. \itemc{Note :} \ttt{MSTU(16)} should not be changed between the generation of an event and the translation of this event record with a \ttt{LUHEPC} call, since this may give an erroneous translation of the event history. \end{subentry} \iteme{MSTU(17) :} (D=0) storage option for \ttt{MSTU(90)} and associated information on $z$ values for heavy-flavour production. \begin{subentry} \iteme{= 0 :} \ttt{MSTU(90)} is reset to zero at each \ttt{LUEXEC} call. This is the appropriate course if \ttt{LUEXEC} is only called once per event, as is normally the case when you do not yourself call \ttt{LUEXEC}. \iteme{= 1 :} you have to reset \ttt{MSTU(90)} to zero yourself before each new event. This is the appropriate course if several \ttt{LUEXEC} calls may appear for one event, i.e. if you call \ttt{LUEXEC} directly. \end{subentry} \iteme{MSTU(19) :} (D=0) advisory warning for unphysical flavour setups in \ttt{LU2ENT}, \ttt{LU3ENT} or \ttt{LU4ENT} calls. \begin{subentry} \iteme{= 0 :} yes. \iteme{= 1 :} no; \ttt{MSTU(19)} is reset to 0 in such a call. \end{subentry} \iteme{MSTU(21) :} (D=2) check on possible errors during program execution. Obviously no guarantee is given that all errors will be caught, but some of the most trivial user-caused errors may be found. \begin{subentry} \iteme{= 0 :} errors do not cause any immediate action, rather the program will try to cope, which may mean e.g. that it runs into an infinite loop. \iteme{= 1 :} parton/particle configurations are checked for possible errors. In case of problem, an exit is made from the misbehaving subprogram, but the generation of the event is continued from there on. For the first \ttt{MSTU(22)} errors a a message is printed; after that no messages appear. \iteme{= 2 :} parton/particle configurations are checked for possible errors. In case of problem, an exit is made from the misbehaving subprogram, and subsequently from \ttt{LUEXEC}. You may then choose to correct the error, and continue the execution by another \ttt{LUEXEC} call. For the first \ttt{MSTU(22)} errors a message is printed, after that the last event is printed and execution is stopped. \end{subentry} \iteme{MSTU(22) :} (D=10) maximum number of errors that are printed. \iteme{MSTU(23) :} (I) count of number of errors experienced to date. \iteme{MSTU(24) :} (R) type of latest error experienced; reason that event was not generated in full. Is reset at each \ttt{LUEXEC} call. \begin{subentry} \iteme{= 0 :} no error experienced. \iteme{= 1 :} have reached end of or are writing outside \ttt{LUJETS} memory. \iteme{= 2 :} unknown flavour code or unphysical combination of codes; may also be caused by erroneous string connection information. \iteme{= 3 :} energy or mass too small or unphysical kinematical variable setup. \iteme{= 4 :} program is caught in an infinite loop. \iteme{= 5 :} momentum, energy or charge was not conserved (even allowing for machine precision errors, see \ttt{PARU(11)}); is evaluated only after event has been generated in full, and does not apply when independent fragmentation without momentum conservation was used. \iteme{= 6 :} error call from outside the fragmentation/decay package (e.g. the $\e^+\e^-$ routines). \iteme{= 7 :} inconsistent particle data input in \ttt{LUUPDA} (\ttt{MUPDA = 2}) or other \ttt{LUUPDA}-related problem. \iteme{= 8 :} problems in more peripheral service routines. \iteme{= 9 :} various other problems. \end{subentry} \iteme{MSTU(25) :} (D=1) printing of warning messages. \begin{subentry} \iteme{= 0 :} no warnings are written. \iteme{= 1 :} first \ttt{MSTU(26)} warnings are printed, thereafter no warnings appear. \end{subentry} \iteme{MSTU(26) :} (D=10) maximum number of warnings that are printed. \iteme{MSTU(27) :} (I) count of number of warnings experienced to date. \iteme{MSTU(28) :} (R) type of latest warning given, with codes paralleling those for \ttt{MSTU(24)}, but of a less serious nature. \iteme{MSTU(31) :} (I) number of \ttt{LUEXEC} calls in present run. \iteme{MSTU(32) :} (I) number of entries stored with \ttt{LUEDIT(-1)} call. \iteme{MSTU(33) :} (I) if set 1 before a \ttt{LUDBRB} call, the \ttt{V} vectors (in the particle range to be rotated/boosted) are set 0 before the rotation/boost. \ttt{MSTU(33)} is set back to 0 in the \ttt{LUDBRB} call. Is inactive in a \ttt{LUROBO} call. \iteme{MSTU(41) - MSTU(63) :} switches for event-analysis routines, see section \ref{ss:evanrout}. \iteme{MSTU(70) :} (D=0) the number of lines consisting only of equal signs (\ttt{======}) that are inserted in the event listing obtained with \ttt{LULIST(1)}, \ttt{LULIST(2)} or \ttt{LULIST(3)}, so as to distinguish different sections of the event record on output. At most 10 such lines can be inserted; see \ttt{MSTU(71) - MSTU(80)}. Is reset at \ttt{LUEDIT} calls with arguments 0--5. \iteme{MSTU(71) - MSTU(80) :} line numbers below which lines consisting only of equal signs (\ttt{======}) are inserted in event listings. Only the first \ttt{MSTU(70)} of the 10 allowed positions are enabled. \iteme{MSTU(90) :} number of heavy-flavour hadrons (i.e. hadrons containing charm or heavier flavours) produced in current event, for which the positions in the event record are stored in \ttt{MSTU(91) - MSTU(98)} and the $z$ values in the fragmentation in \ttt{PARU(91) - PARU(98)}. At most eight values will be stored (normally this is no problem). No $z$ values can be stored for those heavy hadrons produced when a string has so small mass that it collapses to one or two particles, nor for those produced as one of the final two particles in the fragmentation of a string. If \ttt{MSTU(17)=1}, \ttt{MSTU(90)} should be reset to zero by you before each new event, else this is done automatically. \iteme{MSTU(91) - MSTU(98) :} the first \ttt{MSTU(90)} positions will be filled with the line numbers of the heavy-flavour hadrons produced in the current event. See \ttt{MSTU(90)} for additional comments. Note that the information is corrupted by calls to \ttt{LUEDIT} with options 0--5 and 21--23; calls with options 11--15 work, however. \iteme{MSTU(101) - MSTU(118) :} switches related to couplings, see section \ref{ss:coupcons}. \iteme{MSTU(161), MSTU(162) :} information used by event-analysis routines, see section \ref{ss:evanrout}. \iteme{MSTU(181) :} (R) {\tsc{Jetset}} version number. \iteme{MSTU(182) :} (R) {\tsc{Jetset}} subversion number. \iteme{MSTU(183) :} (R) last year of change for {\tsc{Jetset}}. \iteme{MSTU(184) :} (R) last month of change for {\tsc{Jetset}}. \iteme{MSTU(185) :} (R) last day of change for {\tsc{Jetset}}. \iteme{MSTU(186) :} (R) earliest subversion of {\tsc{Pythia}} version 5 with which this {\tsc{Jetset}} subversion can be run. \boxsep \iteme{PARU(1) :}\label{p:PARU} (R) $\pi \approx 3.1415927$. \iteme{PARU(2) :} (R) $2\pi \approx 6.2831854$. \iteme{PARU(3) :} (D=0.1973) conversion factor for GeV$^{-1} \to$ fm or fm$^{-1} \to$ GeV. \iteme{PARU(4) :} (D=5.068) conversion factor for fm $\to$ GeV$^{-1}$ or GeV $\to$ fm$^{-1}$. \iteme{PARU(5) :} (D=0.3894) conversion factor for GeV$^{-2} \to$ mb or mb$^{-1} \to$ GeV$^2$. \iteme{PARU(6) :} (D=2.568) conversion factor for mb $\to$ GeV$^{-2}$ or GeV$^2 \to$ mb$^{-1}$. \iteme{PARU(11) :} (D=0.001) relative error, i.e. nonconservation of momentum and energy divided by total energy, that may be attributable to machine precision problems before a physics error is suspected (see \ttt{MSTU(24)=5}). \iteme{PARU(12) :} (D=0.09 GeV$^2$) effective cut-off in squared mass, below which partons may be recombined to simplify (machine precision limited) kinematics of string fragmentation. \iteme{PARU(13) :} (D=0.01) effective angular cut-off in radians for recombination of partons, used in conjunction with \ttt{PARU(12)}. \iteme{PARU(21) :} (I) contains the total energy $W$ of all first generation jets/particles after a \ttt{LUEXEC} call; to be used by the \ttt{PLU} function for \ttt{I>0}, \ttt{J=} 20--25. \iteme{PARU(41) - PARU(63) :} parameters for event-analysis routines, see section \ref{ss:evanrout}. \iteme{PARU(91) - PARU(98) :} the first \ttt{MSTU(90)} positions will be filled with the fragmentation $z$ values used internally in the generation of heavy-flavour hadrons --- how these are translated into the actual energies and momenta of the observed hadrons is a complicated function of the string configuration. The particle with $z$ value stored in \ttt{PARU(i)} is to be found in line \ttt{MSTU(i)} of the event record. See \ttt{MSTU(90)} and \ttt{MSTU(91) - MSTU(98)} for additional comments. \iteme{PARU(101) - PARU(195) :} various coupling constants and parameters related to couplings, see section \ref{ss:coupcons}. \boxsep \iteme{MSTJ(1) :}\label{p:MSTJ} (D=1) choice of fragmentation scheme. \begin{subentry} \iteme{= 0 :} no jet fragmentation at all. \iteme{= 1 :} string fragmentation according to the Lund model. \iteme{= 2 :} independent fragmentation, according to specification in \ttt{MSTJ(2)} and \ttt{MSTJ(3)}. \end{subentry} \iteme{MSTJ(2) :} (D=3) gluon jet fragmentation scheme in independent fragmentation. \begin{subentry} \iteme{= 1 :} a gluon is assumed to fragment like a random $\d$, $\u$ or $\mathrm{s}$ quark or antiquark. \iteme{= 2 :} as \ttt{=1}, but longitudinal (see \ttt{PARJ(43)}, \ttt{PARJ(44)} and \ttt{PARJ(59)}) and transverse (see \ttt{PARJ(22)}) momentum properties of quark or antiquark substituting for gluon may be separately specified. \iteme{= 3 :} a gluon is assumed to fragment like a pair of a $\d$, $\u$ or $\mathrm{s}$ quark and its antiquark, sharing the gluon energy according to the Altarelli-Parisi splitting function. \iteme{= 4 :} as \ttt{=3}, but longitudinal (see \ttt{PARJ(43)}, \ttt{PARJ(44)} and \ttt{PARJ(59)}) and transverse (see \ttt{PARJ(22)}) momentum properties of quark and antiquark substituting for gluon may be separately specified. \end{subentry} \iteme{MSTJ(3) :} (D=0) energy, momentum and flavour conservation options in independent fragmentation. Whenever momentum conservation is described below, energy and flavour conservation is also implicitly assumed. \begin{subentry} \iteme{= 0 :} no explicit conservation of any kind. \iteme{= 1 :} particles share momentum imbalance compensation according to their energy (roughly equivalent to boosting event to c.m. frame). This is similar to the approach in the Ali et al. program \cite{Ali80}. \iteme{= 2 :} particles share momentum imbalance compensation according to their longitudinal mass with respect to the imbalance direction. \iteme{= 3 :} particles share momentum imbalance compensation equally. \iteme{= 4 :} transverse momenta are compensated separately within each jet, longitudinal momenta are rescaled so that ratio of final jet to initial parton momentum is the same for all the jets of the event. This is similar to the approach in the Hoyer et al. program \cite{Hoy79}. \iteme{= 5 :} only flavour is explicitly conserved. \iteme{= 6 - 10 :} as \ttt{=1 - 5}, except that above several colour singlet systems that followed immediately after each other in the event listing (e.g. $\mathrm{q}\overline{\mathrm{q}}\mathrm{q}\overline{\mathrm{q}}$) were treated as one single system, whereas here they are treated as separate systems. \iteme{= -1 :} independent fragmentation, where also particles moving backwards with respect to the jet direction are kept, and thus the amount of energy and momentum mismatch may be large. \end{subentry} \iteme{MSTJ(11) :} (D=4) choice of longitudinal fragmentation function, i.e. how large a fraction of the energy available a newly-created hadron takes. \begin{subentry} \iteme{= 1 :} the Lund symmetric fragmentation function, see \ttt{PARJ(41) - PARJ(45)}. \iteme{= 2 :} choice of some different forms for each flavour separately, see \ttt{PARJ(51) - PARJ(59)}. \iteme{= 3 :} hybrid scheme, where light flavours are treated with symmetric Lund (\ttt{=1}), but charm and heavier can be separately chosen, e.g. according to the SLAC function (\ttt{=2}). \iteme{= 4 :} the Lund symmetric fragmentation function (\ttt{=1}), for heavy endpoint quarks modified according to the Bowler (Artru--Mennessier, Morris) space--time picture of string evolution, see \ttt{PARJ(46)}. \iteme{= 5 :} as \ttt{=4}, but with possibility to interpolate between Bowler and Lund separately for $\c$, $\b$ and $\t$; see \ttt{PARJ(46) - PARJ(48)}. \end{subentry} \iteme{MSTJ(12) :} (D=2) choice of baryon production model. \begin{subentry} \iteme{= 0 :} no baryon-antibaryon pair production at all; initial diquark treated as a unit. \iteme{= 1 :} diquark-antidiquark pair production allowed; diquark treated as a unit. \iteme{= 2 :} diquark-antidiquark pair production allowed, with possibility for diquark to be split according to the `popcorn' scheme. \iteme{= 3 :} as \ttt{=2}, but additionally the production of first rank baryons may be suppressed by a factor \ttt{PARJ(19)}. \end{subentry} \iteme{MSTJ(13) :} (D=0) generation of transverse momentum for endpoint quark(s) of single quark jet or $\mathrm{q}\overline{\mathrm{q}}$ jet system (in multijet events no endpoint transverse momentum is ever allowed for). \begin{subentry} \iteme{= 0 :} no transverse momentum for endpoint quarks. \iteme{= 1 :} endpoint quarks obtain transverse momenta like ordinary $\mathrm{q}\overline{\mathrm{q}}$ pairs produced in the field (see \ttt{PARJ(21)}); for 2-jet systems the endpoints obtain balancing transverse momenta. \end{subentry} \iteme{MSTJ(14) :} (D=1) treatment of a colour-singlet jet system with a low invariant mass. \begin{subentry} \iteme{= 0 :} no precautions are taken, meaning that problems may occur in \ttt{LUSTRF} (or \ttt{LUINDF}) later on. \iteme{= 1 :} small jet systems are allowed to collapse into two particles or, failing that, one single particle. Normally all small systems are treated this way, starting with the smallest one, but some systems would require more work and are left untreated; they include diquark-antidiquark pairs below the two-particle threshold. \iteme{= -1 :} special option for \ttt{LUPREP} calls, where no precautions are taken (as for \ttt{=0}), but, in addition, no checks are made on the presence of small-mass systems; i.e. \ttt{LUPREP} only rearranges colour strings. \end{subentry} \iteme{MSTJ(15) :} (D=0) production probability for new flavours. \begin{subentry} \iteme{= 0 :} according to standard Lund parametrization, as given by \ttt{PARJ(1) - PARJ(20)}. \iteme{= 1 :} according to probabilities stored in \ttt{PARF(201) - PARF(1960)}; note that no default values exist here, i.e. \ttt{PARF} must be set by you. The \ttt{MSTJ(12)} switch can still be used to set baryon production mode, with the modification that \ttt{MSTJ(12)=2} here allows an arbitrary number of mesons to be produced between a baryon and an antibaryon (since the probability for diquark $\to$ meson $+$ new diquark is assumed independent of prehistory). \end{subentry} \iteme{MSTJ(21) :} (D=2) form of particle decays. \begin{subentry} \iteme{= 0 :} all particle decays are inhibited. \iteme{= 1 :} a particle declared unstable in the \ttt{MDCY} vector, and with decay channels defined, may decay within the region given by \ttt{MSTJ(22)}. A particle may decay into jets, which then fragment further according to the \ttt{MSTJ(1)} value. \iteme{= 2 :} as \ttt{=1}, except that a $\mathrm{q}\overline{\mathrm{q}}$ jet system produced in a decay (e.g. of a $\mathrm{B}$ meson) is always allowed to fragment according to string fragmentation, rather than according to the \ttt{MSTJ(1)} value (this means that momentum, energy and charge are conserved in the decay). \end{subentry} \iteme{MSTJ(22) :} (D=1) cut-off on decay length for a particle that is allowed to decay according to \ttt{MSTJ(21)} and the \ttt{MDCY} value. \begin{subentry} \iteme{= 1 :} a particle declared unstable is also forced to decay. \iteme{= 2 :} a particle is decayed only if its average proper lifetime is larger than \ttt{PARJ(71)}. \iteme{= 3 :} a particle is decayed only if the decay vertex is within a distance \ttt{PARJ(72)} of the origin. \iteme{= 4 :} a particle is decayed only if the decay vertex is within a cylindrical volume with radius \ttt{PARJ(73)} in the $xy$-plane and extent to $\pm$\ttt{PARJ(74)} in the $z$ direction. \end{subentry} \iteme{MSTJ(23) :} (D=1) possibility of having a shower evolving from a $\mathrm{q}\overline{\mathrm{q}}$ pair created as decay products. \begin{subentry} \iteme{= 0 :} never. \iteme{= 1 :} whenever the decay channel matrix-element code is \ttt{MDME(IDC,2)=} 4, 32, 33, 44 or 46, the two first decay products (if they are partons) are allowed to shower, like a colour-singlet subsystem, with maximum virtuality given by the invariant mass of the pair. \end{subentry} \iteme{MSTJ(24) :} (D=2) particle masses. \begin{subentry} \iteme{= 0 :} discrete mass values are used. \iteme{= 1 :} particles registered as having a mass width in the \ttt{PMAS} vector are given a mass according to a truncated Breit--Wigner shape, linear in $m$, eq.~(\ref{dec:BWlin}). \iteme{= 2 :} as \ttt{=1}, but gauge bosons (actually all particles with $|$KF$| \leq 100$) are distributed according to a Breit--Wigner quadratic in $m$, as obtained from propagators. \iteme{= 3 :} as \ttt{=1}, but Breit--Wigner shape is always quadratic in $m$, eq.~(\ref{dec:BWtwo}). \end{subentry} \iteme{MSTJ(25) :} (D=1) inclusion of the $\mathrm{W}^{\pm}$ propagator, in addition to the standard, `infinitely heavy' weak $V-A$ matrix element, in the decay of a $\t$, $\mathrm{l}$ or $\mathrm{h}$ quark, or $\chi$ lepton. \begin{subentry} \iteme{= 0 :} not included. \iteme{= 1 :} included. \end{subentry} \iteme{MSTJ(26) :} (D=2) inclusion of $\mathrm{B}$--$\br{\mathrm{B}}$ mixing in decays. \begin{subentry} \iteme{= 0 :} no. \iteme{= 1 :} yes, with mixing parameters given by \ttt{PARJ(76)} and \ttt{PARJ(77)}. Mixing decays are not specially marked. \iteme{= 2 :} yes, as \ttt{=1}, but a $\mathrm{B}$ ($\br{\mathrm{B}}$) that decays as a $\br{\mathrm{B}}$ ($\mathrm{B}$) is marked as \ttt{K(I,1)=12} rather than the normal \ttt{K(I,1)=11}. \end{subentry} \iteme{MSTJ(27) :} (D=2) possibility for the $\b$ quark to develop a shower in the decay of a top hadron, i.e. $\mathrm{T} \to \mathrm{W}^+ \b \overline{\mathrm{q}}$, where $\overline{\mathrm{q}}$ is a spectator quark. \begin{subentry} \iteme{= 0 :} no, i.e. $\b$ jet is narrow, low-multiplicity. \iteme{= 1 :} the $\b$ is allowed to shower and the $\mathrm{W}$ momentum (in the rest frame of the $\mathrm{T}$) is reduced acccordingly. The $\mathrm{W}$ is therafter assumed to decay isotropically. \iteme{= 2 :} the $\b$ is allowed to shower, as in \ttt{=1}, but the $\mathrm{W}$ decays anisotropically, with the same polarization as in the standard weak decay of option \ttt{=0}. In principle this is better than option \ttt{=1}, but in practice there is no big difference. \end{subentry} \iteme{MSTJ(28) :} (D=0) call to an external $\tau$ decay library. For this option to be meaningful, it is up to you to write the appropriate interface and include that in the routine \ttt{LUTAUD}, as explained in section \ref{ss:JETphysrout}. \begin{subentry} \iteme{= 0 :} not done, i.e. the internal \ttt{LUDECY} treatment is used. \iteme{= 1 :} done whenever the $\tau$ mother particle species can be identified, else the internal \ttt{LUDECY} treatment is used. Normally the mother particle should always be identified, but it is possible for a user to remove event history information or to add extra $\tau$'s directly to the event record, and then the mother is not known. \iteme{= 2 :} always done. \end{subentry} \iteme{MSTJ(40) - MSTJ(50) :} switches for time-like parton showers, see section \ref{ss:showrout}. \iteme{MSTJ(51) :} (D=0) inclusion of Bose--Einstein effects. \begin{subentry} \iteme{= 0 :} no effects included. \iteme{= 1 :} effects included according to an exponential parametrization $C_2(Q) = 1 + $\ttt{PARJ(92)}$\times \exp(-Q/$\ttt{PARJ(93)}$)$, where $C_2(Q)$ represents the ratio of particle production at $Q$ with Bose--Einstein effects to that without, and the relative momentum $Q$ is defined by $Q^2(p_1,p_2) = -(p_1 - p_2)^2 = (p_1 + p_2)^2 - 4m^2$. Particles with width broader than \ttt{PARJ(91)} are assumed to have time to decay before Bose--Einstein effects are to be considered. \iteme{= 2 :} effects included according to a Gaussian parametrization $C_2(Q) = 1 + $\ttt{PARJ(92)}$\times \exp(- (Q/$\ttt{PARJ(93)}$)^2 )$, with notation and comments as above. \end{subentry} \iteme{MSTJ(52) :} (D=3) number of particle species for which Bose--Einstein correlations are to be included, ranged along the chain $\pi^+$, $\pi^-$, $\pi^0$, $\mathrm{K}^+$, $\mathrm{K}^-$, $\mathrm{K}^0_{\mrm{L}}$, $\mathrm{K}^0_{\mrm{S}}$, $\eta$ and $\eta'$. Default corresponds to including all pions ($\pi^+$, $\pi^-$, $\pi^0$), 7 to including all Kaons as well, and 9 is maximum. \iteme{MSTJ(91) :} (I) flag when generating gluon jet with options \ttt{MSTJ(2)=} 2 or 4 (then \ttt{=1}, else \ttt{=0}). \iteme{MSTJ(92) :} (I) flag that a $\mathrm{q}\overline{\mathrm{q}}$ or $\mathrm{g}\g$ pair or a $\mathrm{g}\g\mathrm{g}$ triplet created in \ttt{LUDECY} should be allowed to shower, is 0 if no pair or triplet, is the entry number of the first parton if a pair indeed exists, is the entry number of the first parton, with a $-$ sign, if a triplet indeed exists. \iteme{MSTJ(93) :} (I) switch for \ttt{ULMASS} action. Is reset to 0 in \ttt{ULMASS} call. \begin{subentry} \iteme{= 0 :} ordinary action. \iteme{= 1 :} light ($\d$, $\u$, $\mathrm{s}$, $\c$, $\b$) quark masses are taken from \ttt{PARF(101) - PARF(105)} rather than \ttt{PMAS(1,1) - PMAS(5,1)}. Diquark masses are given as sum of quark masses, without spin splitting term. \iteme{= 2 :} as \ttt{=1}. Additionally the constant terms \ttt{PARF(121)} and \ttt{PARF(122)} are subtracted from quark and diquark masses, respectively. \end{subentry} \iteme{MSTJ(101) - MSTJ(121) :} switches for $\e^+\e^-$ event generation, see section \ref{ss:eeroutines}. \boxsep \iteme{PARJ(1) :}\label{p:PARJ} (D=0.10) is ${\cal P}(\mathrm{q}\q)/{\cal P}(\mathrm{q})$, the suppression of diquark-antidiquark pair production in the colour field, compared with quark--antiquark production. \iteme{PARJ(2) :} (D=0.30) is ${\cal P}(\mathrm{s})/{\cal P}(u)$, the suppression of $\mathrm{s}$ quark pair production in the field compared with $\u$ or $\d$ pair production. \iteme{PARJ(3) :} (D=0.4) is $({\cal P}(\u\mathrm{s})/{\cal P}(\u\d))/({\cal P}(\mathrm{s})/{\cal P}(\d))$, the extra suppression of strange diquark production compared with the normal suppression of strange quarks. \iteme{PARJ(4) :} (D=0.05) is $(1/3){\cal P}(\u\d_1)/{\cal P}(\u\d_0)$, the suppression of spin 1 diquarks compared with spin 0 ones (excluding the factor 3 coming from spin counting). \iteme{PARJ(5) :} (D=0.5) parameter determining relative occurence of baryon production by $BM\br{B}$ and by $B\br{B}$ configurations in the popcorn baryon production model, roughly ${\cal P}(BM\br{B})/({\cal P}(B\br{B})+{\cal P}(BM\br{B})) =$ \ttt{PARJ(5)}$/(0.5+$\ttt{PARJ(5)}$)$. \iteme{PARJ(6) :} (D=0.5) extra suppression for having a $\mathrm{s}\overline{\mathrm{s}}$ pair shared by the $B$ and $\br{B}$ of a $BM\br{B}$ situation. \iteme{PARJ(7) :} (D=0.5) extra suppression for having a strange meson $M$ in a $BM\br{B}$ configuration. \iteme{PARJ(11) - PARJ(17) :} parameters that determine the spin of mesons. \begin{subentry} \iteme{PARJ(11) :} (D=0.5) is the probability that a light meson (containing $\u$ and $\d$ quarks only) has spin 1 (with \ttt{1-PARJ(11)} the probability for spin 0) when formed in fragmentation. \iteme{PARJ(12) :} (D=0.6) is the probability that a strange meson has spin 1. \iteme{PARJ(13) :} (D=0.75) is the probability that a charm or heavier meson has spin 1. \iteme{PARJ(14) :} (D=0.) is the probability that a spin = 0 meson is produced with an orbital angular momentum 1, for a total spin = 1. \iteme{PARJ(15) :} (D=0.) is the probability that a spin = 1 meson is produced with an orbital angular momentum 1, for a total spin = 0. \iteme{PARJ(16) :} (D=0.) is the probability that a spin = 1 meson is produced with an orbital angular momentum 1, for a total spin = 1. \iteme{PARJ(17) :} (D=0.) is the probability that a spin = 1 meson is produced with an orbital angular momentum 1, for a total spin = 2. \itemc{Note :} the end result of the numbers above is that, with \ttt{i} = 11, 12 or 13, depending on flavour content, \\ ${\cal P}(S = 0, L = 0, J = 0) = (1 - \mtt{PARJ(i)}) \times (1 - \mtt{PARJ(14)})$, \\ ${\cal P}(S = 0, L = 1, J = 1) = (1 - \mtt{PARJ(i)}) \times \mtt{PARJ(14)}$, \\ ${\cal P}(S = 1, L = 0, J = 1) = $ \\ \mbox{~}\hfill$\mtt{PARJ(i)} \times (1 - \mtt{PARJ(15)} - \mtt{PARJ(16)} - \mtt{PARJ(17)})$, \\ ${\cal P}(S = 1, L = 1, J = 0) = \mtt{PARJ(i)} \times \mtt{PARJ(15)}$, \\ ${\cal P}(S = 1, L = 1, J = 1) = \mtt{PARJ(i)} \times \mtt{PARJ(16)}$, \\ ${\cal P}(S = 1, L = 1, J = 2) = \mtt{PARJ(i)} \times \mtt{PARJ(17)}$, \\ where $S$ is the quark `true' spin and $J$ is the total spin, usually called the spin $s$ of the meson. \end{subentry} \iteme{PARJ(18) :} (D=1.) is an extra suppression factor multiplying the ordinary {\bf SU(6)} weight for spin $3$/2 baryons, and hence a means to break {\bf SU(6)} in addition to the dynamic breaking implied by \ttt{PARJ(2)}, \ttt{PARJ(3)}, \ttt{PARJ(4)}, \ttt{PARJ(6)} and \ttt{PARJ(7)}. \iteme{PARJ(19) :} (D=1.) extra baryon suppression factor, which multiplies the ordinary diquark-antidiquark production probability for the breakup closest to the endpoint of a string, but leaves other breaks unaffected. Is only used for \ttt{MSTJ(12)=3}. \iteme{PARJ(21) :} (D=0.36 GeV) corresponds to the width $\sigma$ in the Gaussian $p_x$ and $p_y$ transverse momentum distributions for primary hadrons. See also \ttt{PARJ(22) - PARJ(24)}. \iteme{PARJ(22) :} (D=1.) relative increase in transverse momentum in a gluon jet generated with \ttt{MSTJ(2)=} 2 or 4. \iteme{PARJ(23), PARJ(24) :} (D=0.01, 2.) a fraction \ttt{PARJ(23)} of the Gaussian transverse momentum distribution is taken to be a factor \ttt{PARJ(24)} larger than input in \ttt{PARJ(21)}. This gives a simple parametrization of non-Gaussian tails to the Gaussian shape assumed above. \iteme{PARJ(25) :} (D=1.) extra suppression factor for $\eta$ production in fragmentation; if an $\eta$ is rejected a new flavour pair is generated and a new hadron formed. \iteme{PARJ(26) :} (D=0.4) extra suppression factor for $\eta'$ production in fragmentation; if an $\eta'$ is rejected a new flavour pair is generated and a new hadron formed. \iteme{PARJ(31) :} (D=0.1 GeV) gives the remaining $W_+$ below which the generation of a single jet is stopped (it is chosen smaller than a pion mass, so that no hadrons moving in the forward direction are missed). \iteme{PARJ(32) :} (D=1. GeV) is, with quark masses added, used to define the minimum allowable energy of a colour-singlet jet system. \iteme{PARJ(33) - PARJ(34) :} (D=0.8 GeV, 1.5 GeV) are, together with quark masses, used to define the remaining energy below which the fragmentation of a jet system is stopped and two final hadrons formed. \ttt{PARJ(33)} is normally used, except for \ttt{MSTJ(11)=2}, when \ttt{PARJ(34)} is used. \iteme{PARJ(36) :} (D=2.) represents the dependence on the mass of the final quark pair for defining the stopping point of the fragmentation. Is strongly correlated to the choice of \ttt{PARJ(33) - PARJ(35)}. \iteme{PARJ(37) :} (D=0.2) relative width of the smearing of the stopping point energy. \iteme{PARJ(38) - PARJ(39) :} (D=2.5, 0.6) refers to the probability for reverse rapidity ordering of the final two hadrons, according to eq.~(\ref{fr:revord}), where $d_0 = $\ttt{PARJ(39)} for \ttt{MSTJ(11)}$\neq 2$, and $d = $\ttt{PARJ(39)} for \ttt{MSTJ(11)=2}. \iteme{PARJ(41), PARJ(42) :} (D=0.3, 0.58 GeV$^{-2}$) give the $a$ and $b$ parameters of the symmetric Lund fragmentation function for \ttt{MSTJ(11)=}1, 4 and 5 (and \ttt{MSTJ(11)=3} for ordinary hadrons). \iteme{PARJ(43), PARJ(44) :} (D=0.5, 0.9 GeV$^{-2}$) give the $a$ and $b$ parameters as above for the special case of a gluon jet generated with IF and \ttt{MSTJ(2)=} 2 or 4. \iteme{PARJ(45) :} (D=0.5) the amount by which the effective $a$ parameter in the Lund flavour dependent symmetric fragmentation function is assumed to be larger than the normal $a$ when diquarks are produced. More specifically, referring to eq.~(\ref{fr:LSFFlong}), $a_{\alpha} = $\ttt{PARJ(41)} when considering the fragmentation of a quark and = \ttt{PARJ(41) + PARJ(45)} for the fragmentation of a diquark, with corresponding expression for $a_{\beta}$ depending on whether the newly created object is a quark or diquark (for an independent gluon jet generated with \ttt{MSTJ(2)=} 2 or 4, replace \ttt{PARJ(41)} by \ttt{PARJ(43)}). In the popcorn model, a meson created in between the baryon and antibaryon has $a_{\alpha} = a_{\beta} = $\ttt{PARJ(41) + PARJ(45)}. \iteme{PARJ(46) - PARJ(48) :} (D=3*1.) modification of the Lund symmetric fragmentation for heavy endpoint quarks according to the recipe by Bowler, available when \ttt{MSTJ(11)=} 4 or 5 is selected. The shape is given by eq.~(\ref{fr:LSFFBowler}). If \ttt{MSTJ(11)=4} then $r_{\mathrm{Q}} = $\ttt{PARJ(46)} for all flavours, while if \ttt{MSTJ(11)=5} then $r_{\c} = $\ttt{PARJ(46)}, $r_{\b} = $\ttt{PARJ(47)} and $r_{\mathrm{Q}} = $\ttt{PARJ(48)} for $\t$ and heavier. \ttt{PARJ(46) - PARJ(48)} thus provide a possibility to interpolate between the `pure' Bowler shape, $r = 1$, and the normal Lund one, $r = 0$. The additional modifications made in \ttt{PARJ(43) - PARJ(45)} are automatically taken into account, if necessary. \iteme{PARJ(51) - PARJ(58) :} (D=3*0.77, $-0.05$, $-0.005$, 3*$-0.00001$) give a choice of four possible ways to parametrize the fragmentation function for \ttt{MSTJ(11)=2} (and \ttt{MSTJ(11)=3} for charm and heavier). The fragmentation of each flavour KF may be chosen separately; for a diquark the flavour of the heaviest quark is used. With $c = $\ttt{PARJ(50+KF)}, the parametrizations are: \\ $0 \leq c \leq 1$ : Field-Feynman, $f(z) = 1 - c + 3c(1-z)^2$; \\ $-1 \leq c < 0$ : SLAC, $f(z) = 1/(z(1-1/z-(-c)/(1-z))^2)$; \\ $c > 1$ : power peaked at $z=0$, $f(z) = (1-z)^{c-1}$; \\ $c < -1$ : power peaked at $z=1$, $f(z) = z^{-c-1}$. \iteme{PARJ(59) :} (D=1.) replaces \ttt{PARJ(51) - PARJ(53)} for gluon jet generated with \ttt{MSTJ(2)=} 2 or 4. \iteme{PARJ(61) - PARJ(63) :} (D=4.5, 0.7, 0.) parametrizes the energy dependence of the primary multiplicity distribution in phase-space decays. The former two correspond to $c_1$ and $c_2$ of eq.~(\ref{dec:multsel}), while the latter allows a further additive term in the multiplicity specifically for onium decays. \iteme{PARJ(64) :} (0.003 GeV) minimum kinetic energy in decays (safety margin for numerical precision errors). \iteme{PARJ(65) :} (D=0.5 GeV) mass which, in addition to the spectator quark ordiquark mass, is not assumed to partake in the weak decay of a heavy quark in a hadron. \iteme{PARJ(66) :} (D=0.5) relative probability that colour is rearranged when two singlets are to be formed from decay products. Only applies for \ttt{MDME(IDC,2)=} 11--30, i.e. low-mass phase-space decays. \iteme{PARJ(71) :} (D=10 mm) maximum average proper lifetime for particles allowed to decay in the \ttt{MSTJ(22)=2} option. With the default value, $\mathrm{K}_{\mrm{S}}^0$, $\Lambda$, $\Sigma^-$, $\Sigma^+$, $\Xi^-$, $\Xi^0$ and $\Omega^-$ are stable (in addition to those normally taken to be stable), but charm and bottom do still decay. \iteme{PARJ(72) :} (D=1000 mm) maximum distance from the origin at which a decay is allowed to take place in the \ttt{MSTJ(22)=3} option. \iteme{PARJ(73) :} (D=100 mm) maximum cylindrical distance $\rho = \sqrt{x^2 + y^2}$ from the origin at which a decay is allowed to take place in the \ttt{MSTJ(22)=4} option. \iteme{PARJ(74) :} (D=1000 mm) maximum z distance from the origin at which a decay is allowed to take place in the \ttt{MSTJ(22)=4} option. \iteme{PARJ(76) :} (D=0.7) mixing parameter $x_d = \Delta M/\Gamma$ in $\mathrm{B}^0$--$\br{\mathrm{B}}^0$ system. \iteme{PARJ(77) :} (D=10.) mixing parameter $x_s = \Delta M/\Gamma$ in $\mathrm{B}_s^0$--$\br{\mathrm{B}}_s^0$ system. \iteme{PARJ(81) - PARJ(89) :} parameters for time-like parton showers, see section \ref{ss:showrout}. \iteme{PARJ(91) :} (D=0.020 GeV) minimum particle width in \ttt{PMAS(KC,2)}, above which particle decays are assumed to take place before the stage where Bose--Einstein effects are introduced. \iteme{PARJ(92) :} (D=1.) nominal strength of Bose--Einstein effects for $Q = 0$, see \ttt{MSTJ(51)}. This parameter, often denoted $\lambda$, expresses the amount of incoherence in particle production. Due to the simplified picture used for the Bose--Einstein effects, in particular for effects from three nearby identical particles, the actual $\lambda$ of the simulated events may be larger than the input value. \iteme{PARJ(93) :} (D=0.20 GeV) size of the Bose--Einstein effect region in terms of the $Q$ variable, see \ttt{MSTJ(51)}. The more conventional measure, in terms of the radius $R$ of the production volume, is given by $R = \hbar/$\ttt{PARJ(93)}$ \approx 0.2$ fm$\times$GeV/\ttt{PARJ(93)}$ = $\ttt{PARU(3)}/\ttt{PARJ(93)}. \iteme{PARJ(121) - PARJ(171) :} parameters for $\e^+\e^-$ event generation, see section \ref{ss:eeroutines}. \end{entry} \subsection{Couplings} \label{ss:coupcons} In this section we collect information on the two routines for running $\alpha_{\mathrm{s}}$ and $\alpha_{\mathrm{em}}$, and on other couplings of standard and non-standard particles. Although originally begun for {\tsc{Jetset}} applications, this section has rapidly expanded towards the non-standard aspects, and is thus more of interest for {\tsc{Pythia}} applications than for {\tsc{Jetset}} itself. It could therefore equally well have been put somewhere else in this manual. A few couplings indeed appear in the \ttt{PARP} array, see section \ref{ss:PYswitchpar}. \drawbox{ALEM = ULALEM(Q2)}\label{p:ULALEM} \begin{entry} \itemc{Purpose:} to calculate the running electromagnetic coupling constant $\alpha_{\mathrm{em}}$. Expressions used are described in ref. \cite{Kle89}. See \ttt{MSTU(101)}, \ttt{PARU(101)}, \ttt{PARU(103)} and \ttt{PARU(104)}. \iteme{Q2 :} the momentum transfer scale $Q^2$ at which to evaluate $\alpha_{\mathrm{em}}$. \end{entry} \drawbox{ALPS = ULALPS(Q2)}\label{p:ULALPS} \begin{entry} \itemc{Purpose:} to calculate the running strong coupling constant $\alpha_{\mathrm{s}}$. The first- and second-order expressions are given by eqs.~(\ref{ee:aS3j}) and (\ref{ee:aS4j}). See \ttt{MSTU(111) - MSTU(118)} and \ttt{PARU(111) - PARU(118)} for options. \iteme{Q2 :} the momentum transfer scale $Q^2$ at which to evaluate $\alpha_{\mathrm{s}}$. \end{entry} \drawbox{COMMON/LUDAT1/MSTU(200),PARU(200),MSTJ(200),PARJ(200)} \begin{entry} \itemc{Purpose:} to give access to a number of status codes and parameters which regulate the performance of the program as a whole. Here only those related to couplings are described; the main description is found in section \ref{ss:JETswitch}. \boxsep \iteme{MSTU(101) :}\label{p:MSTU101} (D=1) procedure for $\alpha_{\mathrm{em}}$ evaluation in the \ttt{ULALEM} function. \begin{subentry} \iteme{= 0 :} $\alpha_{\mathrm{em}}$ is taken fixed at the value \ttt{PARU(101)}. \iteme{= 1 :} $\alpha_{\mathrm{em}}$ is running with the $Q^2$ scale, taking into account corrections from fermion loops ($\mathrm{e}$, $\mu$, $\tau$, $\d$, $\u$, $\mathrm{s}$, $\c$, $\b$). \iteme{= 2 :} $\alpha_{\mathrm{em}}$ is fixed, but with separate values at low and high $Q^2$. For $Q^2$ below (above) \ttt{PARU(104)} the value \ttt{PARU(101)} (\ttt{PARU(103)}) is used. The former value is then intended for real photon emission, the latter for electroweak physics, e.g. of the $\mathrm{W}/\mathrm{Z}$ gauge bosons. \end{subentry} \iteme{MSTU(111) :} (D=1) order of $\alpha_{\mathrm{s}}$ evaluation in the \ttt{ULALPS} function. Is overwritten in \ttt{LUEEVT}, \ttt{LUONIA} or \ttt{PYINIT} calls with the value desired for the process under study. \begin{subentry} \iteme{= 0 :} $\alpha_{\mathrm{s}}$ is fixed at the value \ttt{PARU(111)}. \iteme{= 1 :} first-order running $\alpha_{\mathrm{s}}$ is used. \iteme{= 2 :} second-order running $\alpha_{\mathrm{s}}$ is used. \end{subentry} \iteme{MSTU(112) :} (D=5) the nominal number of flavours assumed in the $\alpha_{\mathrm{s}}$ expression, with respect to which $\Lambda$ is defined. \iteme{MSTU(113) :} (D=3) minimum number of flavours that may be assumed in $\alpha_{\mathrm{s}}$ expression, see \ttt{MSTU(112)}. \iteme{MSTU(114) :} (D=5) maximum number of flavours that may be assumed in $\alpha_{\mathrm{s}}$ expression, see \ttt{MSTU(112)}. \iteme{MSTU(115) :} (D=0) treatment of $\alpha_{\mathrm{s}}$ singularity for $Q^2 \to 0$. \begin{subentry} \iteme{= 0 :} allow it to diverge like $1/\ln(Q^2/\Lambda^2)$. \iteme{= 1 :} soften the divergence to $1/\ln(1 + Q^2/\Lambda^2)$. \iteme{= 2 :} freeze $Q^2$ evolution below \ttt{PARU(114)}, i.e. the effective argument is $\max(Q^2, $\ttt{PARU(114)}$)$. \end{subentry} \iteme{MSTU(118) :} (I) number of flavours $n_f$ found and used in latest \ttt{ULALPS} call. \boxsep \iteme{PARU(101) :}\label{p:PARU101} (D=0.00729735=1/137.04) $\alpha_{\mathrm{em}}$, the electromagnetic fine structure constant at vanishing momentum transfer. \iteme{PARU(102) :} (D=0.232) $\sin^2 \! \theta_W$, the weak mixing angle of the standard electroweak model. \iteme{PARU(103) :} (D=0.007764=1/128.8) typical $\alpha_{\mathrm{em}}$ in electroweak processes; used for $Q^2 >$\ttt{PARU(104)} in the option \ttt{MSTU(101)=2} of \ttt{ULALEM}. \iteme{PARU(104) :} (D=1 GeV$^2$) dividing line between `low' and `high' $Q^2$ values in the option \ttt{MSTU(101)=2} of \ttt{ULALEM}. \iteme{PARU(105) :} (D=1.16639E-5 GeV$^{-2}$) $G_{\mathrm{F}}$, the Fermi constant of weak interactions. \iteme{PARU(108) :} (I) the $\alpha_{\mathrm{em}}$ value obtained in the latest call to the \ttt{ULALEM} function. \iteme{PARU(111) :} (D=0.20) fix $\alpha_{\mathrm{s}}$ value assumed in \ttt{ULALPS} when \ttt{MSTU(111)=0} (and also in parton showers when $\alpha_{\mathrm{s}}$ is assumed fix there). \iteme{PARU(112) :} (D=0.25 GeV) $\Lambda$ used in running $\alpha_{\mathrm{s}}$ expression in \ttt{ULALPS}. Like \ttt{MSTU(111)}, this value is overwritten by the calling physics routines, and is therefore purely nominal. \iteme{PARU(113) :} (D=1.) the flavour thresholds, for the effective number of flavours $n_f$ to use in the $\alpha_{\mathrm{s}}$ expression, are assumed to sit at $Q^2 = $\ttt{PARU(113)}$\times m_{\mathrm{q}}^2$, where $m_{\mathrm{q}}$ is the quark mass. May be overwritten from the calling physics routine. \iteme{PARU(114) :} (D=4 GeV$^2$) $Q^2$ value below which the $\alpha_{\mathrm{s}}$ value is assumed constant for \ttt{MSTU(115)=2}. \iteme{PARU(115) :} (D=10.) maximum $\alpha_{\mathrm{s}}$ value that \ttt{ULALPS} will ever return; is used as a last resort to avoid singularities. \iteme{PARU(117) :} (I) $\Lambda$ value (associated with \ttt{MSTU(118)} effective flavours) obtained in latest \ttt{ULALPS} call. \iteme{PARU(118) :} (I) $\alpha_{\mathrm{s}}$ value obtained in latest \ttt{ULALPS} call. \iteme{PARU(121) - PARU(130) :} couplings of a new $\mathrm{Z}'^0$; for fermion default values are given by the Standard Model $\mathrm{Z}^0$ values, assuming $\sin^2 \! \theta_W = 0.23$. Note that e.g. the $\mathrm{Z}'^0$ width contains squared couplings, and thus depends quadratically on the values below. \begin{subentry} \iteme{PARU(121), PARU(122) :} (D=-0.693,-1.) vector and axial couplings of down type quarks to $\mathrm{Z}'^0$. \iteme{PARU(123), PARU(124) :} (D=0.387,1.) vector and axial couplings of up type quarks to $\mathrm{Z}'^0$. \iteme{PARU(125), PARU(126) :} (D=-0.08,-1.) vector and axial couplings of leptons to $\mathrm{Z}'^0$. \iteme{PARU(127), PARU(128) :} (D=1.,1.) vector and axial couplings of neutrinos to $\mathrm{Z}'^0$. \iteme{PARU(129) :} (D=1.) the coupling $Z'^0 \to \mathrm{W}^+ \mathrm{W}^-$ is taken to be \ttt{PARU(129)}$\times$(the Standard Model $\mathrm{Z}^0 \to \mathrm{W}^+ \mathrm{W}^-$ coupling)$\times (m_{\mathrm{W}}/m_{\mathrm{Z}'})^2$. This gives a $\mathrm{Z}'^0 \to \mathrm{W}^+ \mathrm{W}^-$ partial width that increases proportionately to the $\mathrm{Z}'^0$ mass. \iteme{PARU(130) :} (D=0.) in the decay chain $\mathrm{Z}'^0 \to \mathrm{W}^+ \mathrm{W}^- \to 4$ fermions, the angular distribution in the $\mathrm{W}$ decays is supposed to be a mixture, with fraction \ttt{1-PARU(130)} corresponding to the same angular distribution between the four final fermions as in $\mathrm{Z}^0 \to \mathrm{W}^+ \mathrm{W}^-$ (mixture of transverse and longitudinal $\mathrm{W}$'s), and fraction \ttt{PARU(130)} corresponding to $\H^0 \to \mathrm{W}^+ \mathrm{W}^-$ the same way (longitudinal $\mathrm{W}$'s). \end{subentry} \iteme{PARU(131) - PARU(136) :} couplings of a new $\mathrm{W}'^{\pm}$; for fermions default values are given by the Standard Model $\mathrm{W}^{\pm}$ values (i.e. $V-A$). Note that e.g. the $\mathrm{W}'^{\pm}$ width contains squared couplings, and thus depends quadratically on the values below. \begin{subentry} \iteme{PARU(131), PARU(132) :} (D=1.,-1.) vector and axial couplings of a quark--antiquark pair to $\mathrm{W}'^{\pm}$; is further multiplied by the ordinary CKM factors. \iteme{PARU(133), PARU(134) :} (D=1.,-1.) vector and axial couplings of a lepton-neutrino pair to $\mathrm{W}'^{\pm}$. \iteme{PARU(135) :} (D=1.) the coupling $\mathrm{W}'^{\pm} \to \mathrm{Z}^0 \mathrm{W}^{\pm}$ is taken to be \ttt{PARU(135)}$\times$(the Standard Model $\mathrm{W}^{\pm} \to \mathrm{Z}^0 \mathrm{W}^{\pm}$ coupling)$\times (m_{\mathrm{W}}/m_{W'})^2$. This gives a $\mathrm{W}'^{\pm} \to \mathrm{Z}^0 \mathrm{W}^{\pm}$ partial width that increases proportionately to the $\mathrm{W}'$ mass. \iteme{PARU(136) :} (D=0.) in the decay chain $\mathrm{W}'^{\pm} \to \mathrm{Z}^0 \mathrm{W}^{\pm} \to 4$ fermions, the angular distribution in the $\mathrm{W}/\mathrm{Z}$ decays is supposed to be a mixture, with fraction \ttt{1-PARU(130)} corresponding to the same angular distribution between the four final fermions as in $\mathrm{W}^{\pm} \to \mathrm{Z}^0 \mathrm{W}^{\pm}$ (mixture of transverse and longitudinal $\mathrm{W}/\mathrm{Z}$'s), and fraction \ttt{PARU(130)} corresponding to $\H^{\pm} \to \mathrm{Z}^0 \mathrm{W}^{\pm}$ the same way (longitudinal $\mathrm{W}/\mathrm{Z}$'s). \end{subentry} \iteme{PARU(141) :} (D=5.) $\tan\beta$ parameter of a two Higgs doublet scenario, i.e. the ratio of vacuum expectation values. This affects mass relations and couplings in the Higgs sector. \iteme{PARU(142) :} (D=1.) the $\mathrm{Z}^0 \to \H^+ \H^-$ coupling is taken to be \ttt{PARU(142)}$\times$(the MSSM $\mathrm{Z}^0 \to \H^+ \H^-$ coupling). \iteme{PARU(143) :} (D=1.) the $\mathrm{Z}'^0 \to \H^+ \H^-$ coupling is taken to be \ttt{PARU(143)}$\times$(the MSSM $\mathrm{Z}^0 \to \H^+ \H^-$ coupling). \iteme{PARU(145) :} (D=1.) quadratically multiplicative factor in the $\mathrm{Z}'^0 \to \mathrm{Z}^0 \H^0$ partial width in left--right-symmetric models, expected to be unity (see \cite{Coc91}). \iteme{PARU(146) :} (D=1.) $\sin(2\alpha)$ parameter, enters quadratically as multiplicative factor in the $\mathrm{W}'^{\pm} \to \mathrm{W}^{\pm} \H^0$ partial width in left--right-symmetric models (see \cite{Coc91}). \iteme{PARU(151) :} (D=1.) multiplicative factor in the $\L_{\mathrm{Q}} \to \mathrm{q} \ell$ squared Yukawa coupling, and thereby in the $\L_{\mathrm{Q}}$ partial width and the $\mathrm{q} \ell \to \L_{\mathrm{Q}}$ and other cross sections. Specifically, $\lambda^2/(4\pi) = $\ttt{PARU(151)}$\times \alpha_{\mathrm{em}}$, i.e. it corresponds to the $k$ factor of \cite{Hew88}. \iteme{PARU(153) :} (D=0.) anomalous magnetic moment of the $\mathrm{W}^{\pm}$; $\eta = \kappa - 1$, where $\eta = 0$ ($\kappa = 1$) is the Standard Model value. \iteme{PARU(155) :} (D=1000. GeV) compositeness scale $\Lambda$. \iteme{PARU(156) :} (D=1.) sign of interference term between standard cross section and composite term ($\eta$ parameter); should be $\pm 1$. \iteme{PARU(157) - PARU(159) :} (D=3*1.) strength of {\bf SU(2)}, {\bf U(1)} and {\bf SU(3)} couplings, respectively, in an excited fermion scenario; cf. $f$, $f'$ and $f_s$ of \cite{Bau90}. \iteme{PARU(161) - PARU(168) :} (D=5*1.,3*0.) multiplicative factors that can be used to modify the default couplings of the $\H^0$ particle in {\tsc{Pythia}}. Note that the factors enter quadratically in the partial widths. The default values correspond to the couplings given in the minimal one-Higgs-doublet Standard Model. \begin{subentry} \iteme{PARU(161) :} $\H^0$ coupling to down type quarks. \iteme{PARU(162) :} $\H^0$ coupling to up type quarks. \iteme{PARU(163) :} $\H^0$ coupling to leptons. \iteme{PARU(164) :} $\H^0$ coupling to $\mathrm{Z}^0$. \iteme{PARU(165) :} $\H^0$ coupling to $\mathrm{W}^{\pm}$. \iteme{PARU(168) :} $\H^0$ coupling to $\H^{\pm}$ in $\gamma\gamma \to \H^0$ loops, in MSSM $\sin(\beta-\alpha)+\cos(2\beta)\sin(\beta+\alpha) / (2\cos^2 \! \theta_W)$. \end{subentry} \iteme{PARU(171) - PARU(178) :} (D=7*1.,0.) multiplicative factors that can be used to modify the default couplings of the $\H'^0$ particle in {\tsc{Pythia}}. Note that the factors enter quadratically in partial widths. The default values for \ttt{PARU(171) - PARU(175)} correspond to the couplings given to $\H^0$ in the minimal one-Higgs-doublet Standard Model, and are therefore not realistic in a two-Higgs-doublet scenario. The default values should be changed appropriately by you. Also the last two default values should be changed; for these the expressions of the minimal supersymmetric Standard Model (MSSM) are given to show parameter normalization. \begin{subentry} \iteme{PARU(171) :} $\H'^0$ coupling to down type quarks. \iteme{PARU(172) :} $\H'^0$ coupling to up type quarks. \iteme{PARU(173) :} $\H'^0$ coupling to leptons. \iteme{PARU(174) :} $\H'^0$ coupling to $\mathrm{Z}^0$. \iteme{PARU(175) :} $\H'^0$ coupling to $W^{\pm}$. \iteme{PARU(176) :} $\H'^0$ coupling to $\H^0 \H^0$, in MSSM $\cos(2\alpha) \cos(\beta+\alpha) - 2 \sin(2\alpha) \sin(\beta+\alpha)$. \iteme{PARU(177) :} $\H'^0$ coupling to $\mathrm{A}^0 \mathrm{A}^0$, in MSSM $\cos(2\beta) \cos(\beta+\alpha)$. \iteme{PARU(178) :} $\H'^0$ coupling to $\H^{\pm}$ in $\gamma \gamma \to \H'^0$ loops, in MSSM $\cos(\beta-\alpha) - \cos(2\beta)\cos(\beta+\alpha) / (2\cos^2 \! \theta_W)$. \end{subentry} \iteme{PARU(181) - PARU(190) :} (D=3*1.,2*0.,2*1.,3*0.) multiplicative factors that can be used to modify the default couplings of the $\mathrm{A}^0$ particle in PYTHIA. Note that the factors enter quadratically in partial widths. The default values for \ttt{PARU(181) - PARU(183)} correspond to the couplings given to $\H^0$ in the minimal one-Higgs-doublet Standard Model, and are therefore not realistic in a two-Higgs-doublet scenario. The default values should be changed appropriately by you. \ttt{PARU(184)} and \ttt{PARU(185)} should be vanishing at the tree level, and are so set; normalization of these couplings agrees with what is used for H and $\H'^0$. Also the other default values should be changed; for these the expressions of the Minimal Supersymmetric Standard Model (MSSM) are given to show parameter normalization. \begin{subentry} \iteme{PARU(181) :} $\mathrm{A}^0$ coupling to down type quarks. \iteme{PARU(182) :} $\mathrm{A}^0$ coupling to up type quarks. \iteme{PARU(183) :} $\mathrm{A}^0$ coupling to leptons. \iteme{PARU(184) :} $\mathrm{A}^0$ coupling to $\mathrm{Z}^0$. \iteme{PARU(185) :} $\mathrm{A}^0$ coupling to $\mathrm{W}^{\pm}$. \iteme{PARU(186) :} $\mathrm{A}^0$ coupling to $\mathrm{Z}^0 \H^0$ (or $\mathrm{Z}^*$ to $\mathrm{A}^0 \H^0$), in MSSM $\cos(\beta-\alpha)$. \iteme{PARU(187) :} $\mathrm{A}^0$ coupling to $\mathrm{Z}^0 \H'^0$ (or $\mathrm{Z}^*$ to $\mathrm{A}^0 \H'^0$), in MSSM $\sin(\beta-\alpha)$. \iteme{PARU(188) :} As \ttt{PARU(186)}, but coupling to $\mathrm{Z}'^0$ rather than $\mathrm{Z}^0$. \iteme{PARU(189) :} As \ttt{PARU(187)}, but coupling to $\mathrm{Z}'^0$ rather than $\mathrm{Z}^0$. \iteme{PARU(190) :} $\mathrm{A}^0$ coupling to $\H^{\pm}$ in $\gamma \gamma \to \mathrm{A}^0$ loops, 0 in MSSM. \end{subentry} \iteme{PARU(191) - PARU(195) :} (D=4*0.,1.) multiplicative factors that can be used to modify the couplings of the $\H^{\pm}$ particle in {\tsc{Pythia}}. Currently only \ttt{PARU(195)} is in use. See above for related comments. \begin{subentry} \iteme{PARU(195) :} $\H^{\pm}$ coupling to $\mathrm{W}^{\pm} \H^0$ (or $\mathrm{W}^{* \pm}$ to $\H^{\pm} \H^0$), in MSSM $\cos(\beta-\alpha)$. \end{subentry} \end{entry} \subsection{Further Parameters and Particle Data} \label{ss:parapartdat} The following common blocks are maybe of a more peripheral interest, with the exception of the \ttt{MDCY} array, which allows a selective inhibiting of particle decays, and masses of not yet discovered particles, such as \ttt{PMAS(6,1)}, the top quark mass. \drawbox{COMMON/LUDAT2/KCHG(500,3),PMAS(500,4),PARF(2000),VCKM(4,4)}% \label{p:LUDAT2} \begin{entry} \itemc{Purpose:} to give access to a number of flavour treatment constants or parameters and particle/parton data. Particle data is stored by compressed code KC rather than by the full KF code. You are reminded that the way to know the KC value is to use the \ttt{LUCOMP} function, i.e. \ttt{KC = LUCOMP(KF)}. \boxsep \iteme{KCHG(KC,1) :}\label{p:KCHG} three times particle/parton charge for compressed code KC. \iteme{KCHG(KC,2) :} colour information for compressed code KC. \begin{subentry} \iteme{= 0 :} colour-singlet particle. \iteme{= 1 :} quark or antidiquark. \iteme{= -1 :} antiquark or diquark. \iteme{= 2 :} gluon. \end{subentry} \iteme{KCHG(KC,3) :} particle/antiparticle distinction for compressed code KC. \begin{subentry} \iteme{= 0 :} the particle is its own antiparticle. \iteme{= 1 :} a nonidentical antiparticle exists. \end{subentry} \boxsep \iteme{PMAS(KC,1) :}\label{p:PMAS} particle/parton mass $m$ (in GeV) for compressed code KC. \iteme{PMAS(KC,2) :} the total width $\Gamma$ (in GeV) of an assumed symmetric Breit--Wigner mass shape for compressed particle code KC. \iteme{PMAS(KC,3) :} the maximum deviation (in GeV) from the \ttt{PMAS(KC,1)} value at which the Breit--Wigner shape above is truncated. (Is used in particle decays, but not in the {\tsc{Pythia}} resonance treatment; cf. the \ttt{CKIN} variables.) \iteme{PMAS(KC,4) :} the average lifetime $\tau$ for compressed particle code KC, with $c \tau$ in mm, i.e. $\tau$ in units of about $3.33 \times 10^{-12}$ s. \boxsep \iteme{PARF(1) - PARF(60) :}\label{p:PARF} give a parametrization of the $\d\overline{\mathrm{d}}$--$\u\overline{\mathrm{u}}$--$\mathrm{s}\overline{\mathrm{s}}$ flavour mixing in production of flavour-diagonal mesons. Numbers are stored in groups of 10, for the six multiplets pseudoscalar, vector, axial vector ($S=0$), scalar, axial vector ($S=1$) and tensor, in this order; see section \ref{sss:mesonprod}. Within each group, the first two numbers determine the fate of a $\d\overline{\mathrm{d}}$ flavour state, the second two that of a $\u\overline{\mathrm{u}}$ one, the next two that of an $\mathrm{s}\overline{\mathrm{s}}$ one, while the last four are unused. Call the numbers of a pair $p_1$ and $p_2$. Then the probability to produce the state with smallest KF code is $1-p_1$, the probability for the middle one is $p_1 - p_2$ and the probability for the one with largest code is $p_2$, i.e. $p_1$ is the probability to produce either of the two `heavier' ones. \iteme{PARF(61) - PARF(80) :} give flavour {\bf SU(6)} weights for the production of a spin 1/2 or spin 3/2 baryon from a given diquark--quark combination. Should not be changed. \iteme{PARF(101) - PARF(108) :} first five contain $\d$, $\u$, $\mathrm{s}$, $\c$ and $\b$ constituent masses, as to be used in mass formulae, and should not be changed. For $\t$, $\mathrm{l}$ and $\mathrm{h}$ masses the current values stored in \ttt{PMAS(6,1) - PMAS(8,1)} are copied in. \iteme{PARF(111), PARF(112) :} (D=0.0, 0.11 GeV) constant terms in the mass formulae for heavy mesons and baryons, respectively (with diquark getting 2/3 of baryon). \iteme{PARF(113), PARF(114) :} (D=0.16,0.048 GeV) factors which, together with Clebsch-Gordan coefficients and quark constituent masses, determine the mass splitting due to spin-spin interactions for heavy mesons and baryons, respectively. The latter factor is also used for the splitting between spin 0 and spin 1 diquarks. \iteme{PARF(115) - PARF(118) :} (D=0.50, 0.45, 0.55, 0.60 GeV), constant mass terms, added to the constituent masses, to get the mass of heavy mesons with orbital angular momentum $L = 1$. The four numbers are for pseudovector mesons with quark spin 0, and for scalar, pseudovector and tensor mesons with quark spin 1, respectively. \iteme{PARF(121), PARF(122) :} (D=0.1, 0.2 GeV) constant terms, which are subtracted for quark and diquark masses, respectively, in defining the allowed phase space in particle decays into partons. \iteme{PARF(201) - PARF(1960) :} (D=1760*0) relative probabilities for flavour production in the \ttt{MSTJ(15)=1} option; to be defined by you before any {\tsc{Jetset}} calls. \\ The index in \ttt{PARF} is of the compressed form \\ $120 + 80 \times$KTAB1$ + 25 \times$KTABS$ + $KTAB3. \\ Here KTAB1 is the old flavour, fixed by preceding fragmentation history, while KTAB3 is the new flavour, to be selected according to the relevant relative probabilities (except for the very last particle, produced when joining two jets, where both KTAB1 and KTAB3 are known). Only the most frequently appearing quarks/diquarks are defined, according to the code $1 = \d$, $2 = \u$, $3 = \mathrm{s}$, $4 = \c$, $5 = \b$, $6 = \t$, $7 = \d\d_1$, $8 = \u\d_0$, $9 = \u\d_1$, $10 = \u\u_1$, $11 = \mathrm{s}\d_0$, $12 = \mathrm{s}\d_1$, $13 = \mathrm{s}\u_0$, $14 = \mathrm{s}\u_1$, $15 = \mathrm{s}\s_1$, $16 = \c\d_0$, $17 = \c\d_1$, $18 = \c\u_0$, $19 = \c\u_1$, $20 = \c\mathrm{s}_0$, $21 = \c\mathrm{s}_1$, $22 = \c\c_1$. These are thus the only possibilities for the new flavour to be produced; for an occasional old flavour not on this list, the ordinary relative flavour production probabilities will be used. \\ Given the initial and final flavour, the intermediate hadron that is produced is almost fixed. (Initial and final diquark here corresponds to `popcorn' production of mesons intermediate between a baryon and an antibaryon). The additional index KTABS gives the spin type of this hadron, with \\ 0 = pseudoscalar meson or $\Lambda$-like spin 1/2 baryon, \\ 1 = vector meson or $\Sigma$-like spin 1/2 baryon, \\ 2 = tensor meson or spin 3/2 baryon. \\ (Some meson multiplets, not frequently produced, are not accessible by this parametrization.) \\ Note that some combinations of KTAB1, KTAB3 and KTABS do not correspond to a physical particle (a $\Lambda$-like baryon must contain three different quark flavours, a $\Sigma$-like one at least two), and that you must see to it that the corresponding \ttt{PARF} entries are vanishing. One additional complication exist when KTAB3 and KTAB1 denote the same flavour content (normally KTAB3$ = $KTAB1, but for diquarks the spin freedom may give KTAB3$ = $KTAB1$\pm 1$): then a flavour neutral meson is to be produced, and here $\d\overline{\mathrm{d}}$, $\u\overline{\mathrm{u}}$ and $\mathrm{s}\overline{\mathrm{s}}$ states mix (heavier flavour states do not, and these are therefore no problem). For these cases the ordinary KTAB3 value gives the total probability to produce either of the mesons possible, while KTAB3$ = $23 gives the relative probability to produce the lightest meson state ($\pi^0$, $\rho^0$, $\a_2^0$), KTAB3$ = $24 relative probability for the middle meson ($\eta$, $\omega$, $\mathrm{f}_2^0$), and KTAB3 = 25 relative probability for the heaviest one ($\eta'$, $\phi$, $f'^0_2$). Note that, for simplicity, these relative probabilities are assumed the same whether initial and final diquark have the same spin or not; the total probability may well be assumed different, however. \\ As a general comment, the sum of \ttt{PARF} values for a given KTAB1 need not be normalized to unity, but rather the program will find the sum of relevant weights and normalize to that. The same goes for the KTAB3$ = $23--25 weights. This makes it straightforward to use one common setup of \ttt{PARF} values and still switch between different \ttt{MSTJ(12)} baryon production modes. \boxsep \iteme{VCKM(I,J) :}\label{p:VCKM} squared matrix elements of the Cabibbo-Kobayashi-Maskawa flavour mixing matrix. \begin{subentry} \iteme{I :} up type generation index, i.e. $1 = \u$, $2 = \c$, $3 = \t$ and $4 = \mathrm{h}$. \iteme{J :} down type generation index, i.e. $1 = \d$, $2 = \mathrm{s}$, $3 = \b$ and $4 = \mathrm{l}$. \end{subentry} \end{entry} \drawbox{COMMON/LUDAT3/MDCY(500,3),MDME(2000,2),BRAT(2000),% KFDP(2000,5)}\label{p:LUDAT3} \begin{entry} \itemc{Purpose:} to give access to particle decay data and parameters. In particular, the \ttt{MDCY(KC,1)} variables may be used to switch on or off the decay of a given particle species, and the \ttt{MDME(IDC,1)} ones to switch on or off an individual decay channel of a particle. For quarks, leptons and gauge bosons, a number of decay channels are included that are not allowed for on-mass-shell particles, see \ttt{MDME(IDC,2)=102}. These channels are not currently used in {\tsc{Jetset}}, but instead find applications in {\tsc{Pythia}}. Particle data is stored by compressed code KC rather than by the full KF code. You are reminded that the way to know the KC value is to use the \ttt{LUCOMP} function, i.e. \ttt{KC = LUCOMP(KF)}. \boxsep \iteme{MDCY(KC,1) :}\label{p:MDCY} switch to tell whether a particle with compressed code KC may be allowed to decay or not. \begin{subentry} \iteme{= 0 :} the particle is not allowed to decay. \iteme{= 1 :} the particle is allowed to decay (if decay information is defined below for the particle). \end{subentry} \iteme{MDCY(KC,2) :} gives the entry point into the decay channel table for compressed particle code KC. Is 0 if no decay channels have been defined. \iteme{MDCY(KC,3) :} gives the total number of decay channels defined for compressed particle code KC, independently of whether they have been assigned a non-vanishing branching ratio or not. Thus the decay channels are found in positions \ttt{MDCY(KC,2)} to \ttt{MDCY(KC,2)+MDCY(KC,3)-1}. \boxsep \iteme{MDME(IDC,1) :}\label{p:MDME} on/off switch for individual decay channel IDC. In addition, a channel may be left selectively open; this has some special applications in {\tsc{Pythia}} which are not currently used in {\tsc{Jetset}}. Effective branching ratios are automatically recalculated for the decay channels left open. Also process cross sections are affected; see section \ref{sss:resdecaycross}. If a particle is allowed to decay by the \ttt{MDCY(KC,1)} value, at least one channel must be left open by you. A list of decay channels with current IDC numbers may be obtained with \ttt{LULIST(12)}. \begin{subentry} \iteme{= -1 :} this is a non-Standard Model decay mode, which by default is assumed not to exist. Normally, this option is used for decays involving fourth generation or $\H^{\pm}$ particles. \iteme{= 0 :} channel is switched off. \iteme{= 1 :} channel is switched on. \iteme{= 2 :} channel is switched on for a particle but off for an antiparticle. It is also on for a particle its own antiparticle, i.e. here it means the same as \ttt{=1}. \iteme{= 3 :} channel is switched on for an antiparticle but off for a particle. It is off for a particle its own antiparticle. \iteme{= 4 :} in the production of a pair of equal or charge conjugate resonances in {\tsc{Pythia}}, say $\H^0 \to \mathrm{W}^+ \mathrm{W}^-$, either one of the resonances is allowed to decay according to this group of channels, but not both. If the two particles of the pair are different, the channel is on. Within {\tsc{Jetset}}, this option only means that the channel is switched off. \iteme{= 5 :} as \ttt{=4}, but an independent group of channels, such that in a pair of equal or charge conjugate resonances the decay of either resonance may be specified independently. If the two particles in the pair are different, the channel is off. Within {\tsc{Jetset}}, this option only means that the channel is switched off. \itemc{Warning:} the two values -1 and 0 may look similar, but in fact are quite different. In neither case the channel so set is generated, but in the latter case the channel still contributes to the total width of a resonance, and thus affects both simulated line shape and the generated cross section when {\tsc{Pythia}} is run. The value 0 is appropriate to a channel we assume exists, even if we are not currently simulating it, while -1 should be used for channels we believe do not exist. In particular, you are warned unwittingly to set fourth generation channels 0 (rather than -1), since by now the support for a fourth generation is small. \itemc{Remark:} all the options above may be freely mixed. The difference, for those cases where both make sense, between using values 2 and 3 and using 4 and 5 is that the latter automatically include charge conjugate states, e.g. $\H^0 \to \mathrm{W}^+ \mathrm{W}^- \to \mathrm{e}^+ \nu_e \d \overline{\mathrm{u}}$ or $\overline{\mathrm{d}} \u \mathrm{e}^- \br{\nu}_e$, but the former only one of them. In calculations of the joint branching ratio, this makes a factor 2 difference. \itemc{Example:} to illustrate the above options, consider the case of a $\mathrm{W}^+ \mathrm{W}^-$ pair. One might then set the following combination of switches for the $\mathrm{W}$:\\ \begin{tabular}{ccl@{\protect\rule{0mm}{\tablinsep}}} channel & value & comment \\ $\u\overline{\mathrm{d}}$ & 1 & allowed for $\mathrm{W}^+$ and $\mathrm{W}^-$ in any combination, \\ $\u\overline{\mathrm{s}}$ & 0 & never produced but contributes to $\mathrm{W}$ width, \\ $\c\overline{\mathrm{d}}$ & 2 & allowed for $\mathrm{W}^+$ only, \\ $\c\overline{\mathrm{s}}$ & 3 & allowed for $\mathrm{W}^-$ only, i.e. properly $\mathrm{W}^- \to \overline{\mathrm{c}}\mathrm{s}$, \\ $\t\overline{\mathrm{b}}$ & 0 & never produced but contributes to $\mathrm{W}$ width \\ & & if the channel is kinematically allowed, \\ $\nu_{\mathrm{e}}\mathrm{e}^+$ & 4 & allowed for one of $\mathrm{W}^+$ or $\mathrm{W}^-$, but not both, \\ $\nu_{\mu}\mu^+$ & 4 & allowed for one of $\mathrm{W}^+$ or $\mathrm{W}^-$, but not both, \\ & & and not in combination with $\nu_{\mathrm{e}}\mathrm{e}^+$, \\ $\nu_{\tau}\tau^+$ & 5 & allowed for the other $\mathrm{W}$, but not both, \\ $\nu_{\chi}\chi^-$ & $-1$ & not produced and does not contribute to $\mathrm{W}$ width.\\ \end{tabular}\\ A $\mathrm{W}^+\mathrm{W}^-$ final state $\u\overline{\mathrm{d}} + \overline{\mathrm{c}}\mathrm{s}$ is allowed, but not its charge conjugate $\overline{\mathrm{u}}\d + \c\overline{\mathrm{s}}$, since the latter decay mode is not allowed for a $\mathrm{W}^+$. The combination $\nu_{\mathrm{e}}\mathrm{e}^+ + \bar{\nu}_{\tau}\tau^-$ is allowed, since the two channels belong to different groups, but not $\nu_{\mathrm{e}}\mathrm{e}^+ + \bar{\nu}_{\mu}\mu^-$, where both belong to the same. Both $\u\overline{\mathrm{d}} + \bar{\nu}_{\tau}\tau^-$ and $\overline{\mathrm{u}}\d + \nu_{\tau}\tau^+$ are allowed, since there is no clash. The full rulebook, for this case, is given by eq.~(\ref{eq:WWallchancombgen}). A term $r_i^2$ means channel $i$ is allowed for $\mathrm{W}^+$ and $\mathrm{W}^-$ simultaneously, a term $r_i r_j$ that channels $i$ and $j$ may be combined, and a term $2 r_i r_j$ that channels $i$ and $j$ may be combined two ways, i.e. that also a charge conjugate combination is allowed. \end{subentry} \iteme{MDME(IDC,2) :} information on special matrix-element treatment for decay channel IDC. In addition to the outline below, special rules apply for the order in which decay products should be given, so that matrix elements and colour flow is properly treated. One such example is the weak matrix elements, which only will be correct if decay products are given in the right order. The program does not police this, so if you introduce channels of your own and use these codes, you should be guided by the existing particle data. \begin{subentry} \iteme{= 0 :} no special matrix-element treatment; partons and particles are copied directly to the event record, with momentum distributed according to phase space. \iteme{= 1 :} $\omega$ and $\phi$ decays into three pions, eq. (\ref{dec:omegphi}). \iteme{= 2 :} $\pi^0$ or $\eta$ Dalitz decay to $\gamma \mathrm{e}^+ \mathrm{e}^-$, eq.~(\ref{dec:Dalitz}). \iteme{= 3 :} used for vector meson decays into two pseudoscalars, to signal non-isotropic decay angle according to eq. (\ref{dec:psvpsps}), where relevant. \iteme{= 4 :} decay of a spin 1 onium resonance to three gluons or to a photon and two gluons, eq.~(\ref{ee:Upsilondec}). The gluons may subsequently develop a shower if \ttt{MSTJ(23)=1}. \iteme{= 11 :} phase-space production of hadrons from the quarks available. \iteme{= 12 :} as \ttt{=11}, but for onia resonances, with the option of modifying the multiplicity distribution separately. \iteme{= 13 :} as \ttt{=11}, but at least three hadrons to be produced (useful when the two-body decays are given explicitly). \iteme{= 14 :} as \ttt{=11}, but at least four hadrons to be produced. \iteme{= 15 :} as \ttt{=11}, but at least five hadrons to be produced. \iteme{= 22 - 30 :} phase-space production of hadrons from the quarks available, with the multiplicity fixed to be \ttt{MDME(IDC,2)-20}, i.e. 2--10. \iteme{= 31 :} two or more quarks and particles are distributed according to phase space. If three or more products, the last product is a spectator quark, i.e. sitting at rest with respect to the decaying hadron. \iteme{= 32 :} a $\mathrm{q}\overline{\mathrm{q}}$ or $\mathrm{g}\g$ pair, distributed according to phase space (in angle), and allowed to develop a shower if \ttt{MSTJ(23)=1}. \iteme{= 33 :} a triplet $\mathrm{q} X \overline{\mathrm{q}}$, where $X$ is either a gluon or a colour-singlet particle; the final particle ($\overline{\mathrm{q}}$) is assumed to sit at rest with respect to the decaying hadron, and the two first particles ($\mathrm{q}$ and $X$) are allowed to develop a shower if \ttt{MSTJ(23)=1}. \iteme{= 41 :} weak decay, where particles are distributed according to phase space, multiplied by a factor from the expected shape of the momentum spectrum of the direct product of the weak decay (the $\nu_{\tau}$ in $\tau$ decay). \iteme{= 42 :} weak decay matrix element for quarks and leptons. Products may be given either in terms of quarks or hadrons, or leptons for some channels. If the spectator system is given in terms of quarks, it is assumed to collapse into one particle from the onset. If the virtual $\mathrm{W}$ decays into quarks, these quarks are converted to particles, according to phase space in the $\mathrm{W}$ rest frame, as in \ttt{=11}. Is intended for $\tau$, charm and bottom. \iteme{= 43 :} as \ttt{=42}, but if the $\mathrm{W}$ decays into quarks, these will either appear as jets or, for small masses, collapse into a one- or two-body system. \iteme{= 44 :} weak decay matrix element for quarks and leptons, where the spectator system may collapse into one particle for a small invariant mass. If the first two decay products are a $\mathrm{q}\overline{\mathrm{q}}'$ pair, they may develop a parton shower if \ttt{MSTJ(23)=1}. Is intended for top and beyond, but largely superseded by the following option. \iteme{= 45 :} weak decay $\mathrm{q} \to \mathrm{W} \mathrm{q}'$ or $\ell \to \mathrm{W} \nu_{\ell}$, where the $\mathrm{W}$ is registered as a decay product and subsequently treated with \ttt{MDME=46}. To distinguish from ordinary $\mathrm{W}$'s on the mass shell, code KF$ = \pm 89$ is used. The virtual $\mathrm{W}$ mass is selected according to the standard weak decay matrix element, times the $\mathrm{W}$ propagator (for \ttt{MSTJ(25)=1}). There may be two or three decay products; if a third this is a spectator taken to sit at rest. The spectator system may collapse into one particle. Is intended for top and beyond. \iteme{= 46 :} $\mathrm{W}$ (KF = 89) decay into $\mathrm{q}\overline{\mathrm{q}}'$ or $\ell \nu_{\ell}$ according to relative probabilities given by couplings (as stored in the \ttt{BRAT} vector) times a dynamical phase-space factor given by the current $\mathrm{W}$ mass. In the decay, the correct $V-A$ angular distribution is generated if the $\mathrm{W}$ origin is known (heavy quark or lepton). This is therefore the second step of a decay with \ttt{MDME=45}. A $\mathrm{q}\overline{\mathrm{q}}'$ pair may subsequently develop a shower if \ttt{MSTJ(23)=1}. \iteme{= 84 - 88 :} map the decay of this particle onto the generic $\c$, $\b$, $\t$, $\mathrm{l}$ or $\mathrm{h}$ decay modes defined for KC$ = $84--88. \iteme{= 48 :} as \ttt{=42}, but require at least three decay products. \iteme{= 101 :} this is not a proper decay channel, but only to be considered as a continuation line for the decay product listing of the immediately preceding channel. Since the \ttt{KFDP} array can contain five decay products per channel, with this code it is possible to define channels with up to ten decay products. It is not allowed to have several continuation lines after each other. \iteme{= 102 :} this is not a proper decay channel for a decaying particle on the mass shell (or nearly so), and is therefore assigned branching ratio 0. For a particle off the mass shell, this decay mode is allowed, however. By including this channel among the others, the switches \ttt{MDME(IDC,1)} may be used to allow or forbid these channels in hard processes, with cross sections to be calculated separately. As an example, $\gamma \to \u \overline{\mathrm{u}}$ is not possible for a massless photon, but is an allowed channel in $\e^+\e^-$ annihilation. \end{subentry} \boxsep \iteme{BRAT(IDC) :}\label{p:BRAT} give branching ratios for the different decay channels. In principle, the sum of branching ratios for a given particle should be unity. Since the program anyway has to calculate the sum of branching ratios left open by the \ttt{MDME(IDC,1)} values and normalize to that, you need not explicitly ensure this normalization, however. (Warnings are printed in \ttt{LUUPDA(2)} calls if the sum is not unity, but this is entirely intended as a help for finding user mistypings.) For decay channels with \ttt{MDME(IDC,2)}$> 80$ the \ttt{BRAT} values are dummy. \boxsep \iteme{KFDP(IDC,J) :}\label{p:KFDP} contain the decay products in the different channels, with five positions \ttt{J=} 1--5 reserved for each channel IDC. The decay products are given following the standard KF code for jets and particles, with 0 for trailing empty positions. Note that the \ttt{MDME(IDC+1,2)=101} option allows you to double the maximum number of decay product in a given channel from 5 to 10, with the five latter products stored \ttt{KFDP(IDC+1,J)}. \end{entry} \drawboxtwo{COMMON/LUDAT4/CHAF(500)}{CHARACTER CHAF*8}\label{p:LUDAT4} \begin{entry} \itemc{Purpose:} to give access to character type variables. \boxsep \iteme{CHAF :}\label{p:CHAF} particle names (excluding charge) according to KC code. \end{entry} \subsection{Miscellaneous Comments} The previous sections have dealt with the subroutine options and variables one at a time. This is certainly important, but for a full use of the capabilities of the program, it is also necessary to understand how to make different pieces work together. This is something that cannot be explained fully in a manual, but must also be learnt by trial and error. This section contains some examples of relationships between subroutines, common blocks and parameters. It also contains comments on issues that did not fit in naturally anywhere else, but still might be useful to have on record. \subsubsection{Interfacing to detector simulation} Very often, the output of the program is to be fed into a subsequent detector simulation program. It therefore becomes necessary to set up an interface between the \ttt{LUJETS} common block and the detector model. Preferrably this should be done via the \ttt{HEPEVT} standard common block, see section \ref{ss:HEPEVT}, but sometimes this may not be convenient. If a \ttt{LUEDIT(2)} call is made, the remaining entries exactly correspond to those an ideal detector could see: all non-decayed particles, with the exception of neutrinos. The translation of momenta should be trivial (if need be, a \ttt{LUROBO} call can be made to rotate the `preferred' $z$ direction to whatever is the longitudinal direction of the detector), and so should the translation of particle codes. In particular, if the detector simulation program also uses the standard Particle Data Group codes, no conversion at all is needed. The problem then is to select which particles are allowed to decay, and how decay vertex information should be used. Several switches regulate which particles are allowed to decay. First, the master switch \ttt{MSTJ(21)} can be used to switch on/off all decays (and it also contains a choice of how fragmentation should be interfaced). Second, a particle must have decay modes defined for it, i.e. the corresponding \ttt{MDCY(KC,2)} and \ttt{MDCY(KC,3)} entries must be non-zero for compressed code \ttt{KC = LUCOMP(KF)}. This is true for all colour neutral particles except the neutrinos, the photon, the proton and the neutron. (This statement is actually not fully correct, since irrelevant `decay modes' with \ttt{MDME(IDC,2)=102} exist in some cases.) Third, the individual switch in \ttt{MDCY(KC,1)} must be on. Of all the particles with decay modes defined, only $\mu^{\pm}$, $\pi^{\pm}$, $\mathrm{K}^{\pm}$ and $\mathrm{K}_{\mrm{L}}^0$ are by default considered stable. Finally, if \ttt{MSTJ(22)} does not have its default value 1, checks are also made on the lifetime of a particle before it is allowed to decay. In the simplest alternative, \ttt{MSTJ(22)=2}, the comparison is based on the average lifetime, or rather $c \tau$, measured in mm. Thus if the limit \ttt{PARJ(71)} is (the default) 10 mm, then decays of $\mathrm{K}_{\mrm{S}}^0$, $\Lambda$, $\Sigma^-$, $\Sigma^+$, $\Xi^-$, $\Xi^0$ and $\Omega^-$ are all switched off, but charm and bottom still decay. No $c \tau$ values below 1 micron are defined. With the two options \ttt{MSTJ(22)=} 3 or 4, a spherical or cylindrical volume is defined around the origin, and all decays taking place inside this volume are ignored. Whenever a particle is in principle allowed to decay, i.e. \ttt{MSTJ(21)} and \ttt{MDCY} on, an proper lifetime is selected once and for all and stored in \ttt{V(I,5)}. The \ttt{K(I,1)} is then also changed to 4. For \ttt{MSTJ(22)=1}, such a particle will also decay, but else it could remain in the event record. It is then possible, at a later stage, to expand the volume inside which decays are allowed, and do a new \ttt{LUEXEC} call to have particles fulfilling the new conditions (but not the old) decay. As a further option, the \ttt{K(I,1)} code may be put to 5, signalling that the particle will definitely decay in the next \ttt{LUEXEC} call, at the vertex position given (by the user) in the \ttt{V} vector. This then allows the {\tsc{Jetset}} decay routines to be used inside a detector simulation program, as follows. For a particle which did not decay before entering the detector, its point of decay is still well defined (in the absence of deflections by electric or magnetic fields), eq.~(\ref{dec:newvertex}). If it interacts before that point, the detector simulation program is left to handle things. If not, the \ttt{V} vector is updated according to the formula above, \ttt{K(I,1)} is set to 5, and \ttt{LUEXEC} is called, to give a set of decay products, that can again be tracked. A further possibility is to force particles to decay into specific decay channels; this may be particularly interesting for charm or bottom physics. The choice of channels left open is determined by the values of the switches \ttt{MDME(IDC,1)} for decay channel IDC (use \ttt{LULIST(12)} to obtain the full listing). One or several channels may be left open; in the latter case effective branching ratios are automatically recalculated without the need for your intervention. It is also possible to differentiate between which channels are left open for particles and which for antiparticles. Lifetimes are not affected by the exclusion of some decay channels. Note that, whereas forced decays can enhance the efficiency for several kinds of studies, it can also introduce unexpected biases, in particular when events may contain several particles with forced decays, cf. section \ref{sss:resdecaycross}. \subsubsection{Parameter values} A non-trivial question is to know which parameter values to use. The default values stored in the program are based on comparisons with LEP $\e^+\e^- \to \mathrm{Z}^0$ data at around 91 GeV \cite{LEP90}, using a parton-shower picture followed by string fragmentation. If fragmentation is indeed an universal phenomenon, as we would like to think, then the same parameters should also apply at other energies and in other processes. The former aspect, at least, seems to be borne out by comparisons with lower-energy PETRA/PEP data. Note, however, that the choice of parameters is intertwined with the choice of perturbative QCD description. If instead matrix elements are used, a best fit to 30 GeV data would require the values \ttt{PARJ(21)=0.40}, \ttt{PARJ(41)=1.0} and \ttt{PARJ(42)=0.7}. With matrix elements one does not expect an energy independence of the parameters, since the effective minimum invariant mass cut-off is then energy dependent, i.e. so is the amount of soft gluon emission effects lumped together with the fragmentation parameters. This is indeed confirmed by the LEP data. A mismatch in the perturbative QCD treatment could also lead to small differences between different processes. It is often said that the string fragmentation model contains a wealth of parameters. This is certainly true, but it must be remembered that most of these deal with flavour properties, and to a large extent factorize from the treatment of the general event shape. In a fit to the latter it is therefore usually enough to consider the parameters of the perturbative QCD treatment, like $\Lambda$ in $\alpha_{\mathrm{s}}$ and a shower cut-off $Q_0$ (or $\alpha_{\mathrm{s}}$ itself and $y_{\mathrm{min}}$, if matrix elements are used), the $a$ and $b$ parameter of the Lund symmetric fragmentation function (\ttt{PARJ(41)} and \ttt{PARJ(42)}) and the width of the transverse momentum distribution ($\sigma = $\ttt{PARJ(21)}). In addition, the $a$ and $b$ parameters are very strongly correlated by the requirement of having the correct average multiplicity, such that in a typical $\chi^2$ plot, the allowed region corresponds to a very narrow but very long valley, stretched diagonally from small ($a$,$b$) pairs to large ones. As to the flavour parameters, these are certainly many more, but most of them are understood qualitiatively within one single framework, that of tunnelling pair production of flavours. Since the use of independent fragmentation has fallen in disrespect, it should be pointed out that the default parameters here are not particularly well tuned to the data. This especially applies if one, in addition to asking for independent fragmentation, also asks for another setup of fragmentation functions, i.e. other than the standard Lund symmetric one. In particular, note that most fits to the popular Peterson et al. (SLAC) heavy-flavour fragmentation function are based on the actual observed spectrum. In a Monte Carlo simulation, one must then start out with something harder, to compensate for the energy lost by initial-state photon radiation and gluon bremsstrahlung. Since independent fragmentation is not collinear safe (i.e, the emission of a collinear gluon changes the properties of the final event), the tuning is strongly dependent on the perturbative QCD treatment chosen. All the parameters needed for a tuning of independent fragmentation are available, however. \subsubsection{Particle properties} The masses of most frequently used particles are taken from tables. For some rare charm and bottom hadrons, and for heavier flavour hadrons, this would be unwieldy, and instead mass formulae are used, based on the quark content. For the known quarks $\d$, $\u$, $\mathrm{s}$, $\c$ and $\b$, the masses used for this purpose are actually the ones stored in positions 101--105 in the \ttt{PARF} vector, rather than the ones found in \ttt{PMAS}. This means that the \ttt{PMAS} masses can be freely changed by you, to modify the masses that appear in the event record, without courting disaster elsewhere (since mass formulae typically contain $1/m$ terms from spin-spin splittings, it is necessary to have the non-zero `constituent' masses here). Thus you should never touch the mass values stored in \ttt{PARF}. For the heavier flavours top, low and high, the current \ttt{PMAS} values are always used. For these flavours, the only individually defined hadrons are the flavour neutral $\eta$, $\Theta$, $\mathrm{h}_1$, $\chi_0$, $\chi_1$ and $\chi_2$ states. A complete change of top mass in the program thus requires changing \ttt{PMAS(6,1)}, \ttt{PMAS(LUCOMP(661),1)}, \ttt{PMAS(LUCOMP(663),1)}, \ttt{PMAS(LUCOMP(665),1)}, \ttt{PMAS(LUCOMP(10661),1)}, \ttt{PMAS(LUCOMP(10663),1)} and \ttt{PMAS(LUCOMP(20663),1)}. Since the latter heavy-flavour-diagonal states are not normally produced in fragmentation, it would be no disaster to forget changing their masses. Most particles have separately defined decay channels. However, there are so many heavy-flavour hadrons with common decay desciptions, that a few `pseudoparticles' have been introduced for generic decays. The most frequently used ones are 84 for charm decays, 85 for bottom decays and 86 for top decays. Instead of a long list of decay channels, several bottom and charm baryons and all top hadrons therefore only have one `decay channel', which is the instruction to make use of the decay data for particle 84/85/86. The spectator quark of the generic decay channels is found as the light quark/diquark of the hadron considered. All the mesons in the bottom and charm sectors are individually defined, as are the $\Lambda_{\c}$ and $\Lambda_{\b}$ states. For top and heavier hadrons, the decay is likely to be so fast that no hadrons at all are produced, but if they are, the generic pseudoparticle approach is a good approximation. The program contains space so that additional new particles may be introduced. Although not completely trivial, this should not be beyond the ability of an ordinary user. Basically, three steps are involved. First, a mechanism of production has to be introduced. This production may well take place in another program, like {\tsc{Pythia}} or some user-written correspondence, where matrix elements are used to select the hard process. In this case the new particle already exists in the \ttt{LUJETS} common block when {\tsc{Jetset}} is called. A new particle, meson, baryon or glueball, may also be a part of the fragmentation process, in which case \ttt{LUKFDI} would have to be suitably modified. The particle might also appear as a decay product from some already existing particle, and then the decay data in \ttt{/LUDAT3/} would have to be expanded; conceivably also \ttt{LUDECY} would be affected. The second step is to teach to program to recognize the new particle. If a KF code in the range 41 to 80 is used, this is automatically taken care of, and in particular the compressed code KC coincides with KF. If a whole sequence of particles is to be introduced, with KF codes paralleling that of ordinary mesons/baryons (a supersymmetric `meson' multiplet, made of a squark plus an antiquark, say), then \ttt{LUCOMP} must be modified to include a mapping from these KF values to currently unused KC ones, like the range 401 - 500. It is the presence of such a mapping that the program uses to accept a given KF code as bona fide. The third and final step is to define the properties of this new particle. Thus particle charge information must be given in \ttt{KCHG}, mass, width and lifetime in \ttt{PMAS}, particle name in \ttt{CHAF}, and decay data in the \ttt{MDCY}, \ttt{MDME}, \ttt{BRAT} and \ttt{KFDP} arrays. This process is most conveniently carried out by using \ttt{LUUPDA(1)} to produce a table of particle data, which can then be modified, and afterwards read back in with \ttt{LUUPDA(2)}. Note that the particle data is to be introduced for the compressed code KC, not for KF proper. \subsection{Examples} A 10 GeV $\u$ quark jet going out along the $+z$ axis is generated with \begin{verbatim} CALL LU1ENT(0,2,10.,0.,0.) \end{verbatim} Note that such a single jet is not required to conserve energy, momentum or flavour. In the generation scheme, particles with negative $p_z$ are produced as well, but these are automatically rejected unless \ttt{MSTJ(3)=-1}. While frequently used in former days, the one-jet generation option is not of much current interest. In e.g. a leptoproduction event a typical situation could be a $\u$ quark going out in the $+z$ direction and a $\u\d_0$ target remnant essentially at rest. (Such a process can be simulated by {\tsc{Pythia}}, but here we illustrate how to do part of it yourself.) The simplest procedure is probably to treat the process in the c.m. frame and boost it to the lab frame afterwards. Hence, if the c.m. energy is 20 GeV and the boost $\beta_z = 0.996$ (corresponding to $x_B = 0.045$), then \begin{verbatim} CALL LU2ENT(0,2,2101,20.) CALL LUROBO(0.,0.,0.,0.,0.996) \end{verbatim} The jets could of course also be defined and allowed to fragment in the lab frame with \begin{verbatim} CALL LU1ENT(-1,2,223.15,0.,0.) CALL LU1ENT(2,12,0.6837,3.1416,0.) CALL LUEXEC \end{verbatim} Note here that the target diquark is required to move in the backwards direction with $E-p_z = m_{\mathrm{p}}(1-x_B)$ to obtain the correct invariant mass for the system. This is, however, only an artefact of using a fixed diquark mass to represent a varying target remnant mass, and is of no importance for the fragmentation. If one wants a nicer-looking event record, it is possible to use the following \begin{verbatim} CALL LU1ENT(-1,2,223.15,0.,0.) MSTU(10)=1 P(2,5)=0.938*(1.-0.045) CALL LU1ENT(2,2101,0.,0.,0.) MSTU(10)=2 CALL LUEXEC \end{verbatim} A 30 GeV $\u\overline{\mathrm{u}}\mathrm{g}$ event with $E_{\u} = 8$ GeV and $E_{\overline{\mathrm{u}}} = 14$ GeV is simulated with \begin{verbatim} CALL LU3ENT(0,2,21,-2,30.,2.*8./30.,2.*14./30.) \end{verbatim} The event will be given in a standard orientation with the $\u$ quark along the $+z$ axis and the $\overline{\mathrm{u}}$ in the $-z, +x$ quadrant. Note that the flavours of the three partons have to be given in the order they are found along a string, if string fragmentation options are to work. Also note that, for 3-jet events, and particularly 4-jet ones, not all setups of kinematical variables $x$ lie within the kinematically allowed regions of phase space. All common block variables can obviously be changed by including the corresponding common block in the user-written main program. Alternatively, the routine \ttt{LUGIVE} can be used to feed in values, with some additional checks on array bounds then performed. A call \begin{verbatim} CALL LUGIVE('MSTJ(21)=3;PMAS(C663,1)=210.;CHAF(401)=funnyino;'// &'PMAS(21,4)=') \end{verbatim} will thus change the value of \ttt{MSTJ(21)} to 3, the value of \ttt{PMAS(LUCOMP(663),1) = PMAS(136,1)} to 210., the value of \ttt{CHAF(401)} to 'funnyino', and print the current value of \ttt{PMAS(21,4)}. Since old and new values of parameters changed are written to output, this may offer a convenient way of documenting non-default values used in a given run. On the other hand, if a variable is changed back and forth frequently, the resulting voluminous output may be undesirable, and a direct usage of the common blocks is then to be recommended (the output can also be switched off, see \ttt{MSTU(13)}). A general rule of thumb is that none of the physics routines (\ttt{LUSTRF}, \ttt{LUINDF}, \ttt{LUDECY}, etc.) should ever be called directly, but only via \ttt{LUEXEC}. This routine may be called repeatedly for one single event. At each call only those entries that are allowed to fragment or decay, and have not yet done so, are treated. Thus \begin{verbatim} CALL LU2ENT(1,1,-1,20.) ! fill 2 jets without fragmenting MSTJ(1)=0 ! inhibit jet fragmentation MSTJ(21)=0 ! inhibit particle decay MDCY(LUCOMP(111),1)=0 ! inhibit pi0 decay CALL LUEXEC ! will not do anything MSTJ(1)=1 ! CALL LUEXEC ! jets will fragment, but no decays MSTJ(21)=2 ! CALL LUEXEC ! particles decay, except pi0 CALL LUEXEC ! nothing new can happen MDCY(LUCOMP(111),1)=1 ! CALL LUEXEC ! pi0:s decay \end{verbatim} A partial exception to the rule above is \ttt{LUSHOW}. Its main application is for internal use by \ttt{LUEEVT}, \ttt{LUDECY}, and \ttt{PYEVNT}, but it can also be directly called by you. Note that a special format for storing colour-flow information in \ttt{K(I,4)} and \ttt{K(I,5)} must then be used. For simple cases, the \ttt{LU2ENT} can be made to take care of that automatically, by calling with the first argument negative. \begin{verbatim} CALL LU2ENT(-1,1,-2,80.) ! store d ubar with colour flow CALL LUSHOW(1,2,80.) ! shower partons CALL LUEXEC ! subsequent fragmentation/decay \end{verbatim} For more complicated configurations, \ttt{LUJOIN} should be used. It is always good practice to list one or a few events during a run to check that the program is working as intended. With \begin{verbatim} CALL LULIST(1) \end{verbatim} all particles will be listed and in addition total charge, momentum and energy of stable entries will be given. For string fragmentation these quantities should be conserved exactly (up to machine precision errors), and the same goes when running independent fragmentation with one of the momentum conservation options. \ttt{LULIST(1)} gives a format that comfortably fits on an 80 column screen, at the price of not giving the complete story. With \ttt{LULIST(2)} a more extensive listing is obtained, and \ttt{LULIST(3)} also gives vertex information. Further options are available, like \ttt{LULIST(12)}, which gives a list of particle data. An event, as stored in the \ttt{LUJETS} common block, will contain the original jets and the whole decay chain, i.e. also particles which subsequently decayed. If parton showers are used, the amount of parton information is also considerable: first the on-shell partons before showers have been considered, then a \ttt{K(I,1)=22} line with total energy of the showering subsystem, after that the complete shower history tree-like structure, starting off with the same initial partons (now off-shell), and finally the end products of the shower rearranged along the string directions. This detailed record is useful in many connections, but if one only wants to retain the final particles, superfluous information may be removed with \ttt{LUEDIT}. Thus e.g. \begin{verbatim} CALL LUEDIT(2) \end{verbatim} will leave you with the final charged and neutral particles, except for neutrinos. The information in \ttt{LUJETS} may be used directly to study an event. Some useful additional quantities derived from these, such as charge and rapidity, may easily be found via the \ttt{KLU} and \ttt{PLU} functions. Thus electric charge \ttt{=PLU(I,6)} (as integer, three times charge \ttt{=KLU(I,6)}) and true rapidity $y$ with respect to the $z$ axis \ttt{= PLU(I,17)}. A number of utility (\ttt{MSTU}, \ttt{PARU}) and physics (\ttt{MSTJ}, \ttt{PARJ}) switches and parameters are available in common block \ttt{LUDAT1}. All of these have sensible default values. Particle data is stored in common blocks \ttt{LUDAT2}, \ttt{LUDAT3} and \ttt{LUDAT4}. Note that the data in the arrays \ttt{KCHG}, \ttt{PMAS}, \ttt{MDCY} and \ttt{CHAF} is not stored by KF code, but by the compressed code KC. This code is not to be learnt by heart, but instead accessed via the conversion function \ttt{LUCOMP}, \ttt{KC = LUCOMP(KF)}. In the particle tables, the following particles are considered stable: the photon, $\mathrm{e}^{\pm}$, $\mu^{\pm}$, $\pi^{\pm}$, $\mathrm{K}^{\pm}$, $\mathrm{K}_{\mrm{L}}^0$, $\mathrm{p}$, $\overline{\mathrm{p}}$, $\mathrm{n}$, $\br{\mathrm{n}}$ and all the neutrinos. It is, however, always possible to inhibit the decay of any given particle by putting the corresponding \ttt{MDCY} value zero or negative, e.g. \ttt{MDCY(LUCOMP(310),1)=0} makes $\mathrm{K}_{\mrm{S}}^0$ and \ttt{MDCY(LUCOMP(3122),1)=0} $\Lambda$ stable. It is also possible to select stability based on the average lifetime (see \ttt{MSTJ(22)}), or based on whether the decay takes place within a given spherical or cylindrical volume around the origin. The Field-Feynman jet model \cite{Fie78} is available in the program by changing the following values: \ttt{MSTJ(1)=2} (independent fragmentation), \ttt{MSTJ(3)=-1} (retain particles with $p_z < 0$; is not mandatory), \ttt{MSTJ(11)=2} (choice of longitudinal fragmentation function, with the $a$ parameter stored in \ttt{PARJ(51) - PARJ(53)}), \ttt{MSTJ(12)=0} (no baryon production), \ttt{MSTJ(13)=1} (give endpoint quarks $p_{\perp}$ as quarks created in the field), \ttt{MSTJ(24)=0} (no mass broadening of resonances), \ttt{PARJ(2)=0.5} ($\mathrm{s}/\u$ ratio for the production of new $\mathrm{q}\overline{\mathrm{q}}$ pairs), \ttt{PARJ(11)=PARJ(12)=0.5} (probability for mesons to have spin 1) and \ttt{PARJ(21)=0.35} (width of Gaussian transverse momentum distribution). In addition only $\d$, $\u$ and $\mathrm{s}$ single quark jets may be generated following the FF recipe. Today the FF `standard jet' concept is probably dead and buried, so the numbers above should more be taken as an example of the flexibility of the program, than as something to apply in practice. A wide range of independent fragmentation options are implemented, to be accessed with the master switch \ttt{MSTJ(1)=2}. In particular, with \ttt{MSTJ(2)=1} a gluon jet is assumed to fragment like a random $\d$, $\overline{\mathrm{d}}$, $\u$, $\overline{\mathrm{u}}$, $\mathrm{s}$ or $\overline{\mathrm{s}}$ jet, while with \ttt{MSTJ(2)=3} the gluon is split into a $\d\overline{\mathrm{d}}$, $\u\overline{\mathrm{u}}$ or $\mathrm{s}\overline{\mathrm{s}}$ pair of jets sharing the energy according to the Altarelli-Parisi splitting function. Whereas energy, momentum and flavour is not explicitly conserved in independent fragmentation, a number of options are available in \ttt{MSTJ(3)} to ensure this `post facto', e.g. \ttt{MSTJ(3)=1} will boost the event to ensure momentum conservation and then (in the c.m. frame) rescale momenta by a common factor to obtain energy conservation, whereas \ttt{MSTJ(3)=4} rather uses a method of stretching the jets in longitudinal momentum along the respective jet axis to keep angles between jets fixed. \clearpage \section{Event Analysis Routines} To describe the complicated geometries encountered in multihadronic events, a number of event measures have been introduced. These measures are intended to provide a global view of the properties of a given event, wherein the full information content of the event is condensed into one or a few numbers. A steady stream of such measures are proposed for different purposes. Many are rather specialized or never catch on, but a few become standards, and are useful to have easy access to. \tsc{Jetset} therefore contains a number of routines that can be called for any event, and that will directly access the event record to extract the required information. In the presentation below, measures have been grouped in three kinds. The first contains simple event shape quantities, such as sphericity and thrust. The second is jet finding algorithms. The third is a mixed bag of particle multiplicities and compositions, factorial moments and energy--energy correlations, put together in a small statistics package. None of the measures presented here are Lorentz invariant. The analysis will be performed in whatever frame the event happens to be given in. It it therefore up to you to decide whether the frame in which events were generated is the right one, or whether events beforehand should be boosted, e.g. to the c.m. frame. You can also decide which particles you want to have affected by the analysis. \subsection{Event Shapes} \label{ss:evshape} In this section we study general event shape variables: sphericity, thrust, Fox-Wolfram moments, and jet masses. These measures are implemented in the routines \ttt{LUSPHE}, \ttt{LUTHRU}, \ttt{LUFOWO} and \ttt{LUJMAS}, respectively. Each event is assumed characterized by the particle four-momentum vectors $p_i = (\mbf{p}_i, E_i)$, with $i = 1, 2, \cdots , n$ an index running over the particles of the event. \subsubsection{Sphericity} The sphericity tensor is defined as \cite{Bjo70} \begin{equation} S^{\alpha \beta} = \frac{\displaystyle \sum_i p^{\alpha}_{i} \, p^{\beta}_{i} } {\displaystyle \sum_i |\mbf{p}_i|^2 } ~, \end{equation} where $\alpha, \beta = 1, 2, 3$ corresponds to the $x$, $y$ and $z$ components. By standard diagonalization of $S^{\alpha \beta}$ one may find three eigenvalues $\lambda_1 \geq \lambda_2 \geq \lambda_3$, with $\lambda_1 + \lambda_2 + \lambda_3 = 1$. The sphericity of the event is then defined as \begin{equation} S = \frac{3}{2} \, (\lambda_2 + \lambda_3) ~, \end{equation} so that $0 \leq S \leq 1$. Sphericity is essentially a measure of the summed $p_{\perp}^2$ with respect to the event axis; a 2-jet event corresponds to $S \approx 0$ and an isotropic event to $S \approx 1$. The aplanarity $A$, with definition $A = \frac{3}{2} \lambda_3$, is constrained to the range $0 \leq A \leq \frac{1}{2}$. It measures the transverse momentum component out of the event plane: a planar event has $A \approx 0$ and an isotropic one $A \approx \frac{1}{2}$. Eigenvectors $\mbf{v}_j$ can be found that correspond to the three eigenvalues $\lambda_j$ of the sphericity tensor. The $\mbf{v}_1$ one is called the sphericity axis (or event axis, if it is clear from the context that sphericity has been used), while the sphericity event plane is spanned by $\mbf{v}_1$ and $\mbf{v}_2$. The sphericity tensor is quadratic in particle momenta. This means that the sphericity value is changed if one particle is split up into two collinear ones which share the original momentum. Thus sphericity is not an infrared safe quantity in QCD perturbation theory. A useful generalization of the sphericity tensor is \begin{equation} S^{(r) \alpha \beta} = \frac{\displaystyle \sum_i |\mbf{p}_i|^{r-2} \, p^{\alpha}_{i} \, p^{\beta}_{i} }{\displaystyle \sum_i |\mbf{p}_i|^r } ~, \end{equation} where $r$ is the power of the momentum dependence. While $r = 2$ thus corresponds to sphericity, $r = 1$ corresponds to linear measures calculable in perturbation theory \cite{Par78}: \begin{equation} S^{(1) \alpha \beta} = \frac{\displaystyle \sum_i \frac{\displaystyle p^{\alpha}_{i} \, p^{\beta}_{i}}{\displaystyle |\mbf{p}_i|} } {\displaystyle \sum_i |\mbf{p}_i| } ~. \end{equation} Eigenvalues and eigenvectors may be defined exactly as before, and therefore also equivalents of $S$ and $A$. These have no standard names; I tend to call them linearized sphericity $S_{\mrm{lin}}$ and linearized aplanarity $A_{\mrm{lin}}$. Quantities that are standard in the literature are instead the combinations \cite{Ell81} \begin{eqnarray} C & = & 3 ( \lambda_1 \lambda_2 + \lambda_1 \lambda_3 + \lambda_2 \lambda_3 ) ~, \\ D & = & 27 \lambda_1 \lambda_2 \lambda_3 ~. \end{eqnarray} Each of these is constrained to be in the range between 0 and 1. Typically, $C$ is used to measure the 3-jet structure and $D$ the 4-jet one, since $C$ is vanishing for a perfect 2-jet event and $D$ is vanishing for a planar event. The $C$ measure is related to the second Fow-Wolfram moment (see below), $C = 1 - H_2$. Noninteger $r$ values may also be used, and corresponding generalized sphericity and aplanarity measures calculated. While perturbative arguments favour $r = 1$, we know that the fragmentation `noise', e.g. from transverse momentum fluctuations, is proportionately larger for low momentum particles, and so $r > 1$ should be better for experimental event axis determinations. The use of too large an $r$ value, on the other hand, puts all the emphasis on a few high-momentum particles, and therefore involves a loss of information. It should then come as no surprise that intermediate $r$ values, of around 1.5, gives the best performance for event axis determinations in 2-jet events, where the theoretical meaning of the event axis is well-defined. The gain in accuracy compared with the more conventional choices $r=2$ or $r=1$ is rather modest, however. \subsubsection{Thrust} The quantity thrust $T$ is defined by \cite{Bra64} \begin{equation} T = \max_{|\mbf{n}| = 1} \, \frac{\displaystyle \sum_i |\mbf{n} \cdot \mbf{p}_i|} {\displaystyle \sum_i |\mbf{p}_i|} ~, \label{em:thrust} \end{equation} and the thrust axis $\mbf{v}_1$ is given by the $\mbf{n}$ vector for which maximum is attained. The allowed range is $1/2 \leq T \leq 1$, with a 2-jet event corresponding to $T \approx 1$ and an isotropic event to $T \approx 1/2$. In passing, we note that this is not the only definition found in the literature. The definitions agree for events studied in the c.m. frame and where all particles are detected. However, a definition like \begin{equation} T = 2 \, \max_{|\mbf{n}| = 1} \, \frac{\displaystyle \left| \sum_i \theta(\mbf{n} \cdot \mbf{p}_i) \, \mbf{p}_i \right| }{\displaystyle \sum_i |\mbf{p}_i|} = 2 \, \max_{\theta_i = 0,1} \, \frac{\displaystyle \left| \sum_i \theta_i \, \mbf{p}_i \right| } {\displaystyle \sum_i |\mbf{p}_i|} \label{em:thrusttwo} \end{equation} (where $\theta(x)$ is the step function, $\theta(x) = 1 $ if $x > 0$, else $\theta(x) = 0$) gives different results than the one above if e.g. only charged particles are detected. It would even be possible to have $T > 1$; to avoid such problems, often an extra fictitious particle is introduced to balance the total momentum \cite{Bra79}. Eq.~(\ref{em:thrust}) may be rewritten as \begin{equation} T = \max_{\epsilon_i = \pm 1} \, \frac{\displaystyle \left| \sum_i \epsilon_i \, \mbf{p}_i \right| } {\displaystyle \sum_i |\mbf{p}_i|} ~. \label{em:thrustthree} \end{equation} (This may also be viewed as applying eq.~(\ref{em:thrusttwo}) to an event with $2 n$ particles, $n$ carrying the momenta $\mbf{p}_i$ and $n$ the momenta $- \mbf{p}_i$, thus automatically balancing the momentum.) To find the thrust value and axis this way, $2^{n-1}$ different possibilities would have to be tested. The reduction by a factor of 2 comes from $T$ being unchanged when all $\epsilon_i \to - \epsilon_i$. Therefore this approach rapidly becomes prohibitive. Other exact methods exist, which `only' require about $4n^2$ combinations to be tried. In the implementation in {\tsc{Jetset}}, a faster alternative method is used, in which the thrust axis is iterated from a starting direction $\mbf{n}^{(0)}$ according to \begin{equation} \mbf{n}^{(j+1)} = \frac{\displaystyle \sum_i \epsilon( \mbf{n}^{(j)} \cdot \mbf{p}_i) \, \mbf{p}_i } {\displaystyle \left| \sum_i \epsilon( \mbf{n}^{(j)} \cdot \mbf{p}_i) \, \mbf{p}_i \right| } \end{equation} (where $\epsilon(x) = 1$ for $x > 0$ and $\epsilon(x) = -1$ for $x < 0$). It is easy to show that the related thrust value will never decrease, $T^{(j+1)} \geq T^{(j)}$. In fact, the method normally converges in 2--4 iterations. Unfortunately, this convergence need not be towards the correct thrust axis but is occasionally only towards a local maximum of the thrust function \cite{Bra79}. We know of no foolproof way around this complication, but the danger of an error may be lowered if several different starting axes $\mbf{n}^{(0)}$ are tried and found to agree. These $\mbf{n}^{(0)}$ are suitably constructed from the $n'$ (by default 4) particles with the largest momenta in the event, and the $2^{n' -1}$ starting directions $\sum_i \epsilon_i \, \mbf{p}_i$ constructed from these are tried in falling order of the corresponding absolute momentum values. When a predetermined number of the starting axes have given convergence towards the same (best) thrust axis this one is accepted. In the plane perpendicular to the thrust axis, a major \cite{MAR79} axis and value may be defined in just the same fashion as thrust, i.e. \begin{equation} M_a = \max_{|\mbf{n}| = 1, \, \mbf{n} \cdot \mbf{v}_1 = 0} \, \frac{\displaystyle \sum_i |\mbf{n} \cdot \mbf{p}_i|} {\displaystyle \sum_i |\mbf{p}_i| } ~. \end{equation} In a plane more efficient methods can be used to find an axis than in three dimensions \cite{Wu79}, but for simplicity we use the same method as above. Finally, a third axis, the minor axis, is defined perpendicular to the thrust and major ones, and a minor value $M_i$ is calculated just as thrust and major. The difference between major and minor is called oblateness, $O = M_a -M_i$. The upper limit on oblateness depends on the thrust value in a not so simple way. In general $O \approx 0$ corresponds to an event symmetrical around the thrust axis and high $O$ to a planar event. As in the case of sphericity, a generalization to arbitrary momentum dependence may easily be obtained, here by replacing the $\mbf{p}_i$ in the formulae above by $|\mbf{p}_i|^{r-1} \, \mbf{p}_i$. This possibility is included, although so far it has not found any experimental use. \subsubsection{Fox-Wolfram moments} The Fox-Wolfram moments $H_l$, $l = 0, 1, 2, \ldots$, are defined by \cite{Fox79} \begin{equation} H_l = \sum_{i,j} \frac{ |\mbf{p}_i| \, |\mbf{p}_j| } {E_{\mrm{vis}}^2} \, P_l (\cos \theta_{ij}) ~, \end{equation} where $\theta_{ij}$ is the opening angle between hadrons $i$ and $j$ and $E_{\mrm{vis}}$ the total visible energy of the event. Note that also autocorrelations, $i = j$, are included. The $P_l(x)$ are the Legendre polynomials, \begin{eqnarray} P_0(x) & = & 1 ~, \nonumber \\ P_1(x) & = & x ~, \nonumber \\ P_2(x) & = & \frac{1}{2} \, (3x^2 -1) ~, \nonumber \\ P_3(x) & = & \frac{1}{2} \, (5x^3 - 3x) ~, \nonumber \\ P_4(x) & = & \frac{1}{8} \, (35x^4 - 30x^2 + 3) ~. \end{eqnarray} To the extent that particle masses may be neglected, $H_0 \equiv 1$. It is customary to normalize the results to $H_0$, i.e. to give $H_{l0} = H_l / H_0$. If momentum is balanced then $H_1 \equiv 0$. 2-jet events tend to give $H_l \approx 1$ for $l$ even and $\approx 0$ for $l$ odd. \subsubsection{Jet masses} The particles of an event may be divided into two classes. For each class a squared invariant mass may be calculated, $M_1^2$ and $M_2^2$. If the assignment of particles is adjusted such that the sum $M_1^2 + M_2^2$ is minimized, the two masses thus obtained are called heavy and light jet mass, $M_{\mrm{H}}$ and $M_{\mrm{L}}$. It has been shown that these quantities are well behaved in perturbation theory \cite{Cla79}. In $\e^+\e^-$ annihilation, the heavy jet mass obtains a contribution from $\mathrm{q}\overline{\mathrm{q}}\mathrm{g}$ 3-jet events, whereas the light mass is non-vanishing only when 4-jet events also are included. In the c.m. frame of an event one has the limits $0 \leq M_{\mrm{H}}^2 \leq E_{\mrm{cm}}^2/3$. In general, the subdivision of particles tends to be into two hemispheres, separated by a plane perpendicular to an event axis. As with thrust, it is time-consuming to find the exact solution. Different approximate strategies may therefore be used. In the program, the sphericity axis is used to perform a fast subdivision into two hemispheres, and thus into two preliminary jets. Thereafter one particle at a time is tested to determine whether the sum $M_1^2 + M_2^2$ would be decreased if that particle were to be assigned to the other jet. The precedure is stopped when no further significant change is obtained. Often the original assignment is retained as it is, i.e. the sphericity axis gives a good separation. This is not a full guarantee, since the program might get stuck in a local mimimum which is not the global one. \subsection{Cluster Finding} \label{ss:clustfind} Global event measures, like sphericity or thrust, can only be used to determine the jet axes for back-to-back 2-jet events. To determine the individual jet axes in events with three or more jets, or with two (main) jets which are not back-to-back, cluster algorithms are customarily used. In these, nearby particles are grouped together into a variable number of clusters. Each cluster has a well-defined direction, given by a suitably weighted average of the constituent particle directions. The cluster algorithms traditionally used in $\e^+\e^-$ and in $\p\p$ physics differ in several respects. The former tend to be spherically symmetric, i.e. have no preferred axis in space, and normally all particles have to be assigned to some jet. The latter pick the beam axis as preferred direction, and make use of variables related to this choice, such as rapidity and transverse momentum; additionally only a fraction of all particles are assigned to jets. This reflects a difference in the underlying physics: in $\p\p$ collisions, the beam remnants found at low transverse momenta are not related to any hard processes, and therefore only provide an unwanted noise to many studies. (Of course, also hard processes may produce particles at low transverse momenta, but at a rate much less than that from soft or semi-hard processes.) Further, the kinematics of hard processes is, to a good approximation, factorized into the hard subprocess itself, which is boost invariant in rapidity, and parton-distribution effects, which determine the overall position of a hard scattering in rapidity. Hence rapidity, azimuthal angle and transverse momentum is a suitable coordinate frame to describe hard processes in. In standard $\e^+\e^-$ annihilation events, on the other hand, the hard process c.m. frame tends to be almost at rest, and the event axis is just about randomly distributed in space, i.e. with no preferred r\^ole for the axis defined by the incoming $\mathrm{e}^{\pm}$. All particle production is initiated by and related to the hard subprocess. Some of the particles may be less easy to associate to a specific jet, but there is no compelling reason to remove any of them from consideration. This does not mean that the separation above is always required. $2\gamma$ events in $\e^+\e^-$ may have a structure with `beam jets' and `hard scattering' jets, for which the $\p\p$ type algorithms might be well suited. Conversely, a heavy particle produced in $\p\p$ collisions could profitably be studied, in its own rest frame, with $\e^+\e^-$ techniques. In the following, particles are only characterized by their three-momenta or, alternatively, their energy and direction of motion. No knowledge is therefore assumed of particle types, or even of mass and charge. Clearly, the more is known, the more sophisticated clustering algorithms can be used. The procedure then also becomes more detector-dependent, and therefore less suitable for general usage. {\tsc{Jetset}} contains two cluster finding routines. \ttt{LUCLUS} is of the $\e^+\e^-$ type and \ttt{LUCELL} of the $\p\p$ one. Each of them allows some variations of the basic scheme. \subsubsection{Cluster finding in an $\e^+\e^-$ type of environment} The usage of cluster algorithms for $\e^+\e^-$ applications started in the late 1970's. A number of different approaches were proposed \cite{Dor81}. Of these, we will here only discuss those based on binary joining. In this kind of approach, initially each final-state particle is considered to be a cluster. Using some distance measure, the two nearest clusters are found. If their distance is smaller than some cut-off value, the two clusters are joined into one. In this new configuration, the two clusters that are now nearest are found and joined, and so on until all clusters are separated by a distance larger than the cut-off. The clusters remaining at the end are often also called jets. Note that, in this approach, each single particle belongs to exactly one cluster. Also note that the resulting jet picture explicitly depends on the cut-off value used. Normally the number of clusters is allowed to vary from event to event, but occasionally it is more useful to have the cluster algorithm find a predetermined number of jets (like 3). The obvious choice for a distance measure is to use squared invariant mass, i.e. for two clusters $i$ and $j$ to define the distance to be \begin{equation} m^2_{ij} = (E_i + E_j)^2 - (\mbf{p}_i + \mbf{p}_j)^2 ~. \end{equation} (Equivalently, one could have used the invariant mass as measure rather than its square; this is just a matter of convenience.) In fact, a number of people (including the author) tried this measure long ago and gave up on it, since it turns out to have severe instability problems. The reason is well understood: in general, particles tend to cluster closer in invariant mass in the region of small momenta. The clustering process therefore tends to start in the center of the event, and only subsequently spread outwards to encompass also the fast particles. Rather than clustering slow particles around the fast ones (where the latter na\"{\i}vely should best represent the jet directions), the invariant mass measure will tend to cluster fast particles around the slow ones. Another instability may be seen by considering the clustering in a simple 2-jet event. By the time that clustering has reached the level of three clusters, the `best' the clustering algorithm can possibly have achieved, in terms of finding three low-mass clusters, is to have one fast cluster around each jet, plus a third slow cluster in the middle. In the last step this third cluster would be joined with one of the fast ones, to produce two final asymmetric clusters: one cluster would contain all the slow particles, also those that visually look like belonging to the opposite jet. A simple binary joining process, with no possiblity to reassign particles between clusters, is therefore not likely to be optimal. The solution adopted by the author \cite{Sjo83} is to reject invariant mass as distance measure. Instead a jet is defined as a collection of particles which have a limited transverse momentum with respect to a common jet axis, and hence also with respect to each other. This picture is clearly inspired by the standard fragmentation picture, e.g. in string fragmentation. A distance measure $d_{ij}$ between two particles (or clusters) with momenta $\mbf{p}_i$ and $\mbf{p}_j$ should thus not depend critically on the longitudinal momenta but only on the relative transverse momentum. A number of such measures were tried, and the one eventually selected is \begin{equation} d_{ij}^2 = \frac{1}{2} \, (|\mbf{p}_i| \, |\mbf{p}_j| - \mbf{p}_i \cdot \mbf{p}_j) \, \frac{4 \, |\mbf{p}_i| \, |\mbf{p}_j|} {(|\mbf{p}_i| + |\mbf{p}_j|)^2} = \frac{4 \, |\mbf{p}_i|^2 \, |\mbf{p}_j|^2 \, \sin^2(\theta_{ij}/2)} {(|\mbf{p}_i| + |\mbf{p}_j|)^2} ~. \end{equation} For small relative angle $\theta_{ij}$, where $2 \sin(\theta_{ij}/2) \approx \sin\theta_{ij}$ and $\cos \theta_{ij} \approx 1$, this measure reduces to \begin{equation} d_{ij} \approx \frac{|\mbf{p}_i \times \mbf{p}_j|} {|\mbf{p}_i + \mbf{p}_j|} ~, \end{equation} where `$\times$' represents the cross product. We therefore see that $d_{ij}$ in this limit has the simple physical interpretation as the transverse momentum of either particle with respect to the direction given by the sum of the two particle momenta. Unlike the approximate expression, however, $d_{ij}$ does not vanish for two back-to-back particles, but is here more related to the invariant mass between them. The basic scheme is of the binary joining type, i.e. initially each particle is assumed to be a cluster by itself. Then the two clusters with smallest relative distance $d_{ij}$ are found and, if $d_{ij} < d_{\mrm{join}}$, with $d_{\mrm{join}}$ some predetermined distance, the two clusters are joined to one, i.e. their momenta are added vectorially to give the momentum of the new cluster. This is repeated until the distance between any two clusters is $> d_{\mrm{join}}$. The number and momenta of these final clusters then represent our reconstruction of the initial jet configuration, and each particle is assigned to one of the clusters. To make this scheme workable, two further ingredients are necessary, however. Firstly, after two clusters have been joined, some particles belonging to the new cluster may actually be closer to another cluster. Hence, after each joining, all particles in the event are reassigned to the closest of the clusters. For particle $i$, this means that the distance $d_{ij}$ to all clusters $j$ in the event has to be evaluated and compared. After all particles have been considered, and only then, are cluster momenta recalculated to take into account any reassignments. To save time, the assignment procedure is not iterated until a stable configuration is reached, but, since all particles are reassigned at each step, such an iteration is effectively taking place in parallel with the cluster joining. Only at the very end, when all $d_{ij} > d_{\mrm{join}}$, is the reassignment procedure iterated to convergence --- still with the possibility to continue the cluster joining if some $d_{ij}$ should drop below $d_{\mrm{join}}$ due to the reassignment. Occasionally, it may occur that the reassignment step leads to an empty cluster, i.e. one to which no particles are assigned. Since such a cluster has a distance $d_{ij} = 0$ to any other cluster, it is automatically removed in the next cluster joining. However, it is possible to run the program in a mode where a minimum number of jets is to be reconstructed. If this minimum is reached with one cluster empty, the particle is found which has largest distance to the cluster it belongs to. That cluster is then split into two, namely the large-distance particle and a remainder. Thereafter the reassignment procedure is continued as before. Secondly, the large multiplicities normally encountered means that, if each particle initially is to be treated as a separate cluster, the program will become very slow. Therefore a smaller number of clusters, for a normal $\e^+\e^-$ event typically 8--12, is constructed as a starting point for the iteration above, as follows. The particle with the highest momentum is found, and thereafter all particles within a distance $d_{ij} < d_{\mrm{init}}$ from it, where $d_{\mrm{init}} \ll d_{\mrm{join}}$. Together these are allowed to form a single cluster. For the remaining particles, not assigned to this cluster, the procedure is iterated, until all particles have been used up. Particles in the central momentum region, $|\mbf{p}| < 2d_{\mrm{init}}$ are treated separately; if their vectorial momentum sum is above $2d_{\mrm{init}}$ they are allowed to form one cluster, otherwise they are left unassigned in the initial configuration. The value of $d_{\mrm{init}}$, as long as reasonably small, has no physical importance, in that the same final cluster configuration will be found as if each particle initially is assumed to be a cluster by itself: the particles clustered at this step are so nearby anyway that they almost inevitably must enter the same jet; additionally the reassignment procedure allows any possible `mistake' to be corrected in later steps of the iteration. Thus the jet reconstruction depends on one single parameter, $d_{\mrm{join}}$, with a clearcut physical meaning of a transverse momentum `jet-resolution power'. Neglecting smearing from fragmentation, $d_{ij}$ between two clusters of equal energy corresponds to half the invariant mass of the two original partons. If one only wishes to reconstruct well separated jets, a large $d_{\mrm{join}}$ should be chosen, while a small $d_{\mrm{join}}$ would allow the separation of close jets, at the cost of sometimes artificially dividing a single jet into two. In particular, $\b$ quark jets may here be a nuisance. The value of $d_{\mrm{join}}$ to use for a fixed jet-resolution power in principle should be independent of the c.m. energy of events, although fragmentation effects may give a contamination of spurious extra jets that increases slowly with $E_{\mrm{cm}}$ for fixed $d_{\mrm{join}}$. Therefore a $d_{\mrm{join}} = 2.5$ GeV was acceptable at PETRA/PEP, while 3--4 GeV may be better for applications at LEP and beyond. This completes the description of the main option of the \ttt{LUCLUS} routine. Variations are possible. One such is to skip the reassignment step, i.e. to make use only of the simple binary joining procedure, without any possibility to reassign particles between jets. (This option is included mainly as a reference, to check how important reassignment really is.) The other main alternative is to replace the distance measure used above with the one used in the JADE algorithm \cite{JAD86}. The JADE cluster algorithm is an attempt to save the invariant mass measure. The distance measure is defined to be \begin{equation} y_{ij} = \frac{2 E_i E_j (1-\cos \theta_{ij})}{E^2_{\mrm{vis}}} ~. \end{equation} Here $E_{\mrm{vis}}$ is the total visible energy of the event. The usage of $E^2_{\mrm{vis}}$ in the denominator rather than $E_{\mrm{cm}}^2$ tends to make the measure less sensitive to detector acceptance corrections; in addition the dimensionless nature of $y_{ij}$ makes it well suited for a comparison of results at different c.m. energies. For the subsequent discussions, this normalization will be irrelevant, however. The $y_{ij}$ measure is very closely related to the squared mass distance measure: the two coincide (up to the difference in normalization) if $m_i = m_j = 0$. However, consider a pair of particles or clusters with non-vanishing individual masses and a fixed pair mass. Then, the larger the net momentum of the pair, the smaller the $y_{ij}$ measure. This somewhat tends to favour clustering of fast particles, and makes the algorithm less unstable than the one based on true invariant mass. The successes of the JADE algorithm are well known: one obtains a very good agreement between the number of partons generated on the matrix-element (or parton-shower) level and the number of clusters reconstructed from the hadrons, such that QCD aspects like the running of $\alpha_{\mathrm{s}}$ can be studied with a minimal dependence on fragmentation effects. Of course, the insensitivity to fragmentation effects depends on the choice of fragmentation model. Fragmentation effects are small in the string model, but not in independent fragmentation scenarios. Although independent fragmentation in itself is not credible, this may be seen as a signal for caution. One should note that the JADE measure still suffers from some of the diseases of the simple mass measure (without reassignments), namely that particles which go in opposite directions may well be joined into the same cluster. Therefore, while the JADE algorithm is a good way to find the number of jets, it is inferior to the standard $d_{ij}$ measure for a determination of jet directions and energies \cite{Bet92}. The $d_{ij}$ measure also gives narrower jets, which agree better with the visual impression of jet structure. Recently, the `Durham algorithm' has been introduced \cite{Cat91}, which works as the JADE one but with a distance measure \begin{equation} \tilde{y}_{ij} = \frac{2 \min(E_i^2,E_j^2) (1-\cos \theta_{ij})} {E^2_{cm}} ~. \end{equation} Like the $d_{ij}$ measure, this is a transverse momentum, but $\tilde{y}_{ij}$ has the geometrical interpretation as the transverse momentum of the softer particle with respect to the direction of the harder one, while $d_{ij}$ is the transverse momentum of either particle with respect to the common direction given by the momentum vector sum. The two definitions agree when one cluster is much softer than the other, so the soft gluon exponentiation proven for the Durham measure also holds for the $d_{ij}$ one. The main difference therefore is that the standard \ttt{LUCLUS} option allows reassignments, while the Durham algorithm does not. The latter is therefore more easily calculable on the perturbative parton level. This point is sometimes overstressed, and one could give counterexamples why reassignments in fact may bring better agreement with the underlying perturbative level. In particular, without reassignments, one will make the recombination that seems the `best' in the current step, even when that forces you to make `worse' choices in subsequent steps. With reassignments, it is possible to correct for mistakes due to the too local sensitivity of a simple binary joining scheme. \subsubsection{Cluster finding in a $\p\p$ type of environment} The \ttt{LUCELL} cluster finding routines is of the kind pioneered by UA1 \cite{UA183}, and commonly used in $\p\p$ physics. It is based on a choice of pseudorapidity $\eta$, azimuthal angle $\varphi$ and transverse momentum $p_{\perp}$ as the fundamental coordinates. This choice is discussed in the introduction to cluster finding above, with the proviso that the theoretically preferred true rapidity has to be replaced by pseudorapidity, to make contact with the real-life detector coordinate system. A fix detector grid is assumed, with the pseudorapidity range $|\eta| < \eta_{\mathrm{max}}$ and the full azimuthal range each divided into a number of equally large bins, giving a rectangular grid. The particles of an event impinge on this detector grid. For each cell in ($\eta$,$\varphi$) space, the transverse momentum which enters that cell is summed up to give a total cell $E_{\perp}$ flow. Clearly the model remains very primitive in a number of respects, compared with a real detector. There is no magnetic field allowed for, i.e. also charged particles move in straight tracks. The dimensions of the detector are not specified; hence the positions of the primary vertex and any secondary vertices are neglected when determining which cell a particle belongs to. The rest mass of particles is not taken into account, i.e. what is used is really $p_{\perp} = \sqrt{p_x^2 + p_y^2}$, while in a real detector some particles would decay or annihilate, and then deposit additional amounts of energy. To take into account the energy resolution of the detector, it is possible to smear the $E_{\perp}$ contents, bin by bin. This is done according to a Gaussian, with a width assumed proportional to the $\sqrt{E_{\perp}}$ of the bin. The Gaussian is cut off at zero and at some predetermined multiple of the unsmeared $E_{\perp}$, by default twice it. Alternatively, the smearing may be performed in $E$ rather than in $E_{\perp}$. To find the $E$, it is assumed that the full energy of a cell is situated at its center, so that one can translate back and forth with $E = E_{\perp} \cosh\eta_{\mrm{center}}$. The cell with largest $E_{\perp}$ is taken as a jet initiator if its $E_{\perp}$ is above some threshold. A candidate jet is defined to consist of all cells which are within some given radius $R$ in the ($\eta$,$\varphi$) plane, i.e. which have $(\eta - \eta_{\mrm{initiator}})^2 + (\varphi - \varphi_{\mrm{initiator}})^2 < R^2$. Coordinates are always given with respect to the center of the cell. If the summed $E_{\perp}$ of the jet is above the required minimum jet energy, the candidate jet is accepted, and all its cells are removed from further consideration. If not, the candidate is rejected. The sequence is now repeated with the remaining cell of highest $E_{\perp}$, and so on until no single cell fulfills the jet initiator condition. The number of jets reconstruced can thus vary from none to a maximum given by purely geometrical considerations, i.e. how many circles of radius $R$ are needed to cover the allowed ($\eta$,$\varphi$) plane. Normally only a fraction of the particles are assigned to jets. One could consider to iterate the jet assignment process, using the $E_{\perp}$-weighted center of a jet to draw a new cirle of radius $R$. In the current algorithm there is no such iteration step. For an ideal jet assignment it would also be necessary to improve the treatment when two jet circles partially overlap. A final technical note. A natural implementation of a cell finding algorithm is based on having a two-dimensional array of $E_{\perp}$ values, with dimensions to match the detector grid. Very often most of the cells would then be empty, in particular for low-multiplicity events in fine-grained calorimeters. Our implementation is somewhat atypical, since cells are only reserved space (contents and position) when they are shown to be non-empty. This means that all non-empty cells have to be looped over to find which are within the required distance $R$ of a potential jet initiator. The algorithm is therefore faster than the ordinary kind if the average cell occupancy is low, but slower if it is high. \subsection{Event Statistics} All the event-analysis routines above are defined on an event-by-event basis. Once found, the quantities are about equally often used to define inclusive distributions as to select specific classes of events for continued study. For instance, the thrust routine might be used either to find the inclusive $T$ distribution or to select events with $T < 0.9$. Other measures, although still defined for the individual event, only make sense to discuss in terms of averages over many events. A small set of such measures is found in \ttt{LUTABU}. This routine has to be called once after each event to accumulate statistics, and once in the end to print the final tables. Of course, among the wealth of possibilities imaginable, the ones collected here are only a small sample, selected because the author at some point has found a use for them himself. \subsubsection{Multiplicities} Three options are available to collect information on multiplicities in events. One gives the flavour content of the final state in hard interaction processes, e.g. the relative composition of $\d\overline{\mathrm{d}} / \u\overline{\mathrm{u}} / \mathrm{s}\overline{\mathrm{s}} / \c\overline{\mathrm{c}} / \b\overline{\mathrm{b}}$ in $\e^+\e^-$ annihilation events. Additionally it gives the total parton multiplicity distribution at the end of parton showering. Another gives the inclusive rate of all the different particles produced in events, either as intermediate resonances or as final-state particles. The number is subdivided into particles produced from fragmentation (primary particles) and those produced in decays (secondary particles). The third option tabulates the rate of exclusive final states, after all allowed decays have occurred. Since only events with up to 8 final-state particles are analyzed, this is clearly not intended for the study of complete high-energy events. Rather the main application is for an analysis of the decay modes of a single particle. For instance, the decay data for $\mathrm{D}$ mesons is given in terms of channels that also contain unstable particles, such as $\rho$ and $\eta$, which decay further. Therefore a given final state may receive contributions from several tabulated decay channels; e.g. $\mathrm{K} \pi \pi$ from $\mathrm{K}^* \pi$ and $\mathrm{K} \rho$, and so on. \subsubsection{Energy-Energy Correlation} The Energy-Energy Correlation is defined by \cite{Bas78} \begin{equation} \mrm{EEC}(\theta) = \sum_{i < j} \frac{2 E_i E_j}{E_{\mrm{vis}}^2} \, \delta(\theta - \theta_{ij}) ~, \end{equation} and its Asymmetry by \begin{equation} \mrm{EECA}(\theta) = \mrm{EEC}(\pi - \theta) - \mrm{EEC}(\theta) ~. \end{equation} Here $\theta_{ij}$ is the opening angle between the two particles $i$ and $j$, with energies $E_i$ and $E_j$. In principle, normalization should be to $E_{\mrm{cm}}$, but if not all particles are detected it is convenient to normalize to the total visible energy $E_{\mrm{vis}}$. Taking into account the autocorrelation term $i = j$, the total $\mrm{EEC}$ in an event then is unity. The $\delta$ function peak is smeared out by the finite bin width $\Delta \theta$ in the histogram, i.e., it is replaced by a contribution $1 / \Delta \theta$ to the bin which contains $\theta_{ij}$. The formulae above refer to an individual event, and are to be averaged over all events to suppress statistical fluctuations, and obtain smooth functions of $\theta$. \subsubsection{Factorial moments} Factorial moments may be used to search for intermittency in events \cite{Bia86}. The whole field has been much studied in recent years, and a host of different measures have been proposed. We only implement one of the original prescriptions. To calculate the factorial moments, the full rapidity (or pseudorapidity) and azimuthal ranges are subdivided into bins of successively smaller size, and the multiplicity distributions in bins is studied. The program calculates pseudorapidity with respect to the $z$ axis; if desired, one could first find an event axis, e.g. the sphericity or thrust axis, and subsequently rotate the event to align this axis with the $z$ direction. The full rapidity range $|y| < y_{\mathrm{max}}$ (or pseudorapidity range $|\eta| < \eta_{\mathrm{max}}$) and azimuthal range $0 < \varphi < 2\pi$ are subdivided into $m_y$ and $m_{\varphi}$ equally large bins. In fact, the whole analysis is performed thrice: once with $m_{\varphi}=1$ and the $y$ (or $\eta$) range gradually divided into 1, 2, 4, 8, 16, 32, 64, 128, 256 and 512 bins, once with $m_y = 1$ and the $\varphi$ range subdivided as above, and finally once with $m_y = m_{\varphi}$ according to the same binary sequence. Given the multiplicity $n_j$ in bin $j$, the $i$:th factorial moment is defined by \begin{equation} F_i = (m_y m_{\varphi})^{i-1} \, \sum_j \frac{n_j(n_j-1)\cdots(n_j-i+1)}{n(n-1)\cdots(n-i+1)} ~. \end{equation} Here $n = \sum_j n_j$ is the total multiplicity of the event within the allowed $y$ (or $\eta$) limits. The calculation is performed for the second through the fifth moments, i.e. $F_2$ through $F_5$. The $F_i$ as given here are defined for the individual event, and have to be averaged over many events to give a reasonably smooth behaviour. If particle production is uniform and uncorrelated according to Poissonian statistics, one expects $\langle F_i \rangle \equiv 1$ for all moments and all bin sizes. If, on the other hand, particles are locally clustered, factorial moments should increase when bins are made smaller, down to the characteristic dimensions of the clustering. \subsection{Routines and Common Block Variables} \label{ss:evanrout} The six routines \ttt{LUSPHE}, \ttt{LUTHRU}, \ttt{LUCLUS}, \ttt{LUCELL}, \ttt{LUJMAS} and \ttt{LUFOWO} give you the possibility to find some global event shape properties. The routine \ttt{LUTABU} performs a statistical analysis of a number of different quantities like particle content, factorial moments and the energy--energy correlation. Note that, by default, all remaining partons/particles except neutrinos are used in the analysis. Neutrinos may be included with \ttt{MSTU(41)=1}. Also note that axes determined are stored in \ttt{LUJETS}, but are not proper four-vectors and, as a general rule (with some exceptions), should therefore not be rotated or boosted. \drawbox{CALL LUSPHE(SPH,APL)}\label{p:LUSPHE} \begin{entry} \itemc{Purpose:} to diagonalize the momentum tensor, i.e. find the eigenvalues $\lambda_1 > \lambda_2 > \lambda_3$, with sum unity, and the corresponding eigenvectors. \\ Momentum power dependence is given by \ttt{PARU(41)}; default corresponds to sphericity, \ttt{PARU(41)=1.} gives measures linear in momenta. Which particles (or partons) are used in the analysis is determined by the \ttt{MSTU(41)} value. \iteme{SPH :} $\frac{3}{2} (\lambda_2 + \lambda_3)$, i.e. sphericity (for \ttt{PARU(41)=2.}). \begin{subentry} \iteme{= -1. :} analysis not performed because event contained less than two particles (or two exactly back-to-back particles, in which case the two transverse directions would be undefined). \end{subentry} \iteme{APL :} $\frac{3}{2} \lambda_3$, i.e. aplanarity (for \ttt{PARU(41)=2.}). \begin{subentry} \iteme{= -1. :} as \ttt{SPH=-1.}. \end{subentry} \itemc{Remark:} the lines \ttt{N+1} through \ttt{N+3} (\ttt{N-2} through \ttt{N} for \ttt{MSTU(43)=2}) in \ttt{LUJETS} will, after a call, contain the following information: \\ \ttt{K(N+i,1) =} 31; \\ \ttt{K(N+i,2) =} 95; \\ \ttt{K(N+i,3) :} $i$, the axis number, $i=1,2,3$; \\ \ttt{K(N+i,4), K(N+i,5) =} 0; \\ \ttt{P(N+i,1) - P(N+i,3) :} the $i$'th eigenvector, $x$, $y$ and $z$ components; \\ \ttt{P(N+i,4) :} $\lambda_i$, the $i$'th eigenvalue; \\ \ttt{P(N+i,5) =} 0; \\ \ttt{V(N+i,1) - V(N+i,5) =} 0. \\ Also, the number of particles used in the analysis is given in \ttt{MSTU(62)}. \end{entry} \drawbox{CALL LUTHRU(THR,OBL)}\label{p:LUTHRU} \begin{entry} \itemc{Purpose:} to find the thrust, major and minor axes and corresponding projected momentum quantities, in particular thrust and oblateness. The performance of the program is affected by \ttt{MSTU(44)}, \ttt{MSTU(45)}, \ttt{PARU(42)} and \ttt{PARU(48)}. In particular, \ttt{PARU(42)} gives the momentum dependence, with the default value \ttt{=1.} corresponding to linear dependence. Which particles (or partons) are used in the analysis is determined by the \ttt{MSTU(41)} value. \iteme{THR :} thrust (for \ttt{PARU(42)=1.}). \begin{subentry} \iteme{= -1. :} analysis not performed because event contained less than two particles. \iteme{= -2. :} remaining space in \ttt{LUJETS} (partly used as working area) not large enough to allow analysis. \end{subentry} \iteme{OBL :} oblateness (for \ttt{PARU(42)=1.}). \begin{subentry} \iteme{= -1., -2. :} as for \ttt{THR}. \end{subentry} \itemc{Remark:} the lines \ttt{N+1} through \ttt{N+3} (\ttt{N-2} through \ttt{N} for \ttt{MSTU(43)=2}) in \ttt{LUJETS} will, after a call, contain the following information: \\ \ttt{K(N+i,1) =} 31; \\ \ttt{K(N+i,2) =} 96; \\ \ttt{K(N+i,3) :} $i$, the axis number, $i=1,2,3$; \\ \ttt{K(N+i,4), K(N+i,5) =} 0; \\ \ttt{P(N+i,1) - P(N+i,3) :} the thrust, major and minor axis, respectively, for $i = 1, 2$ and 3; \\ \ttt{P(N+i,4) :} corresponding thrust, major and minor value; \\ \ttt{P(N+i,5) =} 0; \\ \ttt{V(N+i,1) - V(N+i,5) =} 0. \\ Also, the number of particles used in the analysis is given in \ttt{MSTU(62)}. \end{entry} \drawbox{CALL LUCLUS(NJET)}\label{p:LUCLUS} \begin{entry} \itemc{Purpose:} to reconstruct an arbitrary number of jets using a cluster analysis method based on particle momenta. \\ Two different distance measures are available, see section \ref{ss:clustfind}. The choice is controlled by \ttt{MSTU(46)}. The distance scale $d_{\mrm{join}}$, above which two clusters may not be joined, is normally given by \ttt{PARU(44)}. In general, $d_{\mrm{join}}$ may be varied to describe different `jet-resolution powers'; the default value, 2.5 GeV, is fairly well suited for $\e^+\e^-$ physics at 30--40 GeV. With the alternative mass distance measure, \ttt{PARU(44)} can be used to set the absolute maximum cluster mass, or \ttt{PARU(45)} to set the scaled one, i.e. in $y = m^2/E_{\mrm{cm}}^2$, where $E_{\mrm{cm}}$ is the total invariant mass of the particles being considered. \\ It is possible to continue the cluster search from the configuration already found, with a new higher $d_{\mrm{join}}$ scale, by selecting \ttt{MSTU(48)} properly. In \ttt{MSTU(47)} one can also require a minimum number of jets to be reconstructed; combined with an artificially large $d_{\mrm{join}}$ this can be used to reconstruct a predetermined number of jets. \\ Which particles (or partons) are used in the analysis is determined by the \ttt{MSTU(41)} value, whereas assumptions about particle masses is given by \ttt{MSTU(42)}. The parameters \ttt{PARU(43)} and \ttt{PARU(48)} regulate more technical details (for events at high energies and large multiplicities, however, the choice of a larger \ttt{PARU(43)} may be necessary to obtain reasonable reconstruction times). \iteme{NJET :} the number of clusters reconstructed. \begin{subentry} \iteme{= -1 :} analysis not performed because event contained less than \ttt{MSTU(47)} (normally 1) particles, or analysis failed to reconstruct the requested number of jets. \iteme{= -2 :} remaining space in \ttt{LUJETS} (partly used as working area) not large enough to allow analysis. \end{subentry} \itemc{Remark:} if the analysis does not fail, further information is found in \ttt{MSTU(61) - MSTU(63)} and \ttt{PARU(61) - PARU(63)}. In particular, \ttt{PARU(61)} contains the invariant mass for the system analyzed, i.e. the number used in determining the denominator of $y = m^2/E_{\mrm{cm}}^2$. \ttt{PARU(62)} gives the generalized thrust, i.e. the sum of (absolute values of) cluster momenta divided by the sum of particle momenta (roughly the same as multicity). \ttt{PARU(63)} gives the minimum distance $d$ (in $p_{\perp}$ or $m$) between two clusters in the final cluster configuration, 0 in case of only one cluster. \\ Further, the lines \ttt{N+1} through \ttt{N+NJET} (\ttt{N-NJET+1} through \ttt{N} for \ttt{MSTU(43)=2}) in \ttt{LUJETS} will, after a call, contain the following information: \\ \ttt{K(N+i,1) =} 31; \\ \ttt{K(N+i,2) =} 97; \\ \ttt{K(N+i,3) :} $i$, the jet number, with the jets arranged in falling order of absolute momentum; \\ \ttt{K(N+i,4) :} the number of particles assigned to jet $i$; \\ \ttt{K(N+i,5) =} 0; \\ \ttt{P(N+i,1) - P(N+i,5) :} momentum, energy and invariant mass of jet $i$; \\ \ttt{V(N+i,1) - V(N+i,5) =} 0. \\ Also, for a particle which was used in the analysis, \ttt{K(I,4)}$ = i$, where \ttt{I} is the particle number and $i$ the number of the jet it has ben assigned to. Undecayed particles not used then have \ttt{K(I,4)=0}. An exception is made for lines with \ttt{K(I,1)=3} (which anyhow are not normally interesting for cluster search), where the colour-flow information stored in \ttt{K(I,4)} is left intact. \end{entry} \drawbox{CALL LUCELL(NJET)}\label{p:LUCELL} \begin{entry} \itemc{Purpose:} to provide a simpler cluster routine more in line with what is currently used in the study of high-$p_{\perp}$ collider events. \\ A detector is assumed to stretch in pseudorapidity between \ttt{-PARU(51)} and \ttt{+PARU(51)} and be segmented in \ttt{MSTU(51)} equally large $\eta$ (pseudorapidity) bins and \ttt{MSTU(52)} $\varphi$ (azimuthal) bins. Transverse energy $E_{\perp}$ for undecayed entries are summed up in each bin. For \ttt{MSTU(53)} non-zero, the energy is smeared by calorimetric resolution effects, cell by cell. This is done according to a Gaussian distribution; if \ttt{MSTU(53)=1} the standard deviation for the $E_{\perp}$ is \ttt{PARU(55)}$\times \sqrt{E_{\perp}}$, if \ttt{MSTU(53)=2} the standard deviation for the $E$ is \ttt{PARU(55)}$\times \sqrt{E}$, $E_{\perp}$ and $E$ expressed in GeV. The Gaussian is cut off at 0 and at a factor \ttt{PARU(56)} times the correct $E_{\perp}$ or $E$. Cells with an $E_{\perp}$ below a given threshold \ttt{PARU(58)} are removed from further consideration; by default \ttt{PARU(58)=0.} and thus all cells are kept. \\ All bins with $E_{\perp} > $\ttt{PARU(52)} are taken to be possible initiators of jets, and are tried in falling $E_{\perp}$ sequence to check whether the total $E_{\perp}$ summed over cells no more distant than \ttt{PARU(54)} in $\sqrt{(\Delta\eta)^2 + (\Delta\varphi)^2}$ exceeds \ttt{PARU(53)}. If so, these cells define one jet, and are removed from further consideration. Contrary to \ttt{LUCLUS}, not all particles need be assigned to jets. Which particles (or partons) are used in the analysis is determined by the \ttt{MSTU(41)} value. \iteme{NJET :} the number of jets reconstructed (may be 0). \begin{subentry} \iteme{= -2 :} remaining space in \ttt{LUJETS} (partly used as working area) not large enough to allow analysis. \end{subentry} \itemc{Remark:} the lines \ttt{N+1} through \ttt{N+NJET} (\ttt{N-NJET+1} through \ttt{N} for \ttt{MSTU(43)=2}) in \ttt{LUJETS} will, after a call, contain the following information: \\ \ttt{K(N+i,1) =} 31; \\ \ttt{K(N+i,2) =} 98; \\ \ttt{K(N+i,3) :} $i$, the jet number, with the jets arranged in falling order in $E_{\perp}$; \\ \ttt{K(N+i,4) :} the number of particles assigned to jet $i$; \\ \ttt{K(N+i,5) =} 0; \\ \ttt{V(N+i,1) - V(N+i,5) =} 0. \\ Further, for \ttt{MSTU(54)=1} \\ \ttt{P(N+i,1), P(N+i,2) =} position in $\eta$ and $\varphi$ of the center of the jet initiator cell, i.e. geometrical center of jet; \\ \ttt{P(N+i,3), P(N+i,4) =} position in $\eta$ and $\varphi$ of the $E_{\perp}$-weighted center of the jet, i.e. the center of gravity of the jet; \\ \ttt{P(N+i,5) =} sum $E_{\perp}$ of the jet; \\ while for \ttt{MSTU(54)=2} \\ \ttt{P(N+i,1) - P(N+i,5) :} the jet momentum vector, constructed from the summed $E_{\perp}$ and the $\eta$ and $\varphi$ of the $E_{\perp}$-weighted center of the jet as \\ $(p_x, p_y, p_z, E, m)=E_{\perp} (\cos\varphi, \sin\varphi, \sinh\eta, \cosh\eta, 0)$; \\ and for \ttt{MSTU(54)=3} \\ \ttt{P(N+i,1) - P(N+i,5) :} the jet momentum vector, constructed by adding vectorially the momentum of each cell assigned to the jet, assuming that all the $E_{\perp}$ was deposited at the center of the cell, and with the jet mass in \ttt{P(N+i,5)} calculated from the summed $E$ and $\mbf{p}$ as $m^2 = E^2 - p_x^2 - p_y^2 - p_z^2$. \\ Also, the number of particles used in the analysis is given in \ttt{MSTU(62)}, and the number of cells hit in \ttt{MSTU(63)}. \end{entry} \drawbox{CALL LUJMAS(PMH,PML)}\label{p:LUJMAS} \begin{entry} \itemc{Purpose:} to reconstruct high and low jet mass of an event. A simplified algorithm is used, wherein a preliminary division of the event into two hemispheres is done transversely to the sphericity axis. Then one particle at a time is reassigned to the other hemisphere if that reduces the sum of squares of the two jet masses, $m_{\mrm{H}}^2 + m_{\mrm{L}}^2$. The procedure is stopped when no further significant change (see \ttt{PARU(48)}) is obtained. Often, the original assignment is retained as it is. Which particles (or partons) used in the analysis is determined by the \ttt{MSTU(41)} value, whereas assumptions about particle masses is given by \ttt{MSTU(42)}. \iteme{PMH :} heavy jet mass (in GeV). \begin{subentry} \iteme{= -2. :} remaining space in \ttt{LUJETS} (partly used as working area) not large enough to allow analysis. \end{subentry} \iteme{PML :} light jet mass (in GeV). \begin{subentry} \iteme{= -2. :} as for \ttt{PMH=-2.}. \end{subentry} \itemc{Remark:} After a successful call, \ttt{MSTU(62)} contains the number of particles used in the analysis, and \ttt{PARU(61)} the invariant mass of the system analyzed. The latter number is helpful in constructing scaled jet masses. \end{entry} \drawbox{CALL LUFOWO(H10,H20,H30,H40)}\label{p:LUFOWO} \begin{entry} \itemc{Purpose:} to do an event analysis in terms of the Fox-Wolfram moments. The moments $H_i$ are normalized to the lowest one, $H_0$. Which particles (or partons) are used in the analysis is determined by the \ttt{MSTU(41)} value. \iteme{H10 :} $H_1/H_0$. Is $=0$ if momentum is balanced. \iteme{H20 :} $H_2/H_0$. \iteme{H30 :} $H_3/H_0$. \iteme{H40 :} $H_4/H_0$. \itemc{Remark:} the number of particles used in the analysis is given in \ttt{MSTU(62)}. \end{entry} \drawbox{CALL LUTABU(MTABU)}\label{p:LUTABU} \begin{entry} \itemc{Purpose:} to provide a number of event-analysis options which can be be used on each new event, with accumulated statistics to be written out on request. When errors are quoted, these refer to the uncertainty in the average value for the event sample as a whole, rather than to the spread of the individual events, i.e. errors decrease like one over the square root of the number of events analyzed. For a correct use of \ttt{LUTABU}, it is not permissible to freely mix generation and analysis of different classes of events, since only one set of statistics counters exists. A single run may still contain sequential `subruns', between which statistics is reset. Whenever an event is analyzed, the number of particles/partons used is given in \ttt{MSTU(62)}. \iteme{MTABU :} determines which action is to be taken. Generally, a last digit equal to 0 indicates that the statistics counters for this option is to be reset; since the counters are reset (by \ttt{DATA} statements) at the beginning of a run, this is not used normally. Last digit 1 leads to an analysis of current event with respect to the desired properties. Note that the resulting action may depend on how the event generated has been rotated, boosted or edited before this call. The statistics accumulated is output in tabular form with last digit 2, while it is dumped in the \ttt{LUJETS} common block for last digit 3. The latter option may be useful for interfacing to graphics output. \begin{subentry} \iteme{= 10 :} statistics on parton multiplicity is reset. \iteme{= 11 :} the parton content of the current event is analyzed, classified according to the flavour content of the hard interaction and the total number of partons. The flavour content is assumed given in \ttt{MSTU(161)} and \ttt{MSTU(162)}; these are automatically set e.g. in \ttt{LUEEVT} and \ttt{PYEVNT} calls. \iteme{= 12 :} gives a table on parton multiplicity distribution. \iteme{= 13 :} stores the parton multiplicity distribution of events in \ttt{/LUJETS/}, using the following format: \\ \ttt{N =} total number of different channels found; \\ \ttt{K(I,1) =} 32; \\ \ttt{K(I,2) =} 99; \\ \ttt{K(I,3), K(I,4) =} the two flavours of the flavour content; \\ \ttt{K(I,5) =} total number of events found with flavour content of \ttt{K(I,3)} and \ttt{K(I,4)}; \\ \ttt{P(I,1) - P(I,5) =} relative probability to find given flavour content and a total of 1, 2, 3, 4 or 5 partons, respectively; \\ \ttt{V(I,1) - V(I,5) =} relative probability to find given flavour content and a total of 6--7, 8--10, 11--15, 16--25 or above 25 partons, respectively. \\ In addition, \ttt{MSTU(3)=1} and \\ \ttt{K(N+1,1) =} 32; \\ \ttt{K(N+1,2) =} 99; \\ \ttt{K(N+1,5) =} number of events analyzed. \iteme{= 20 :} statistics on particle content is reset. \iteme{= 21 :} the particle/parton content of the current event is analyzed, also for particles which have subsequently decayed and partons which have fragmented (unless this has been made impossible by a preceding \ttt{LUEDIT} call). Particles are subdivided into primary and secondary ones, the main principle being that primary particles are those produced in the fragmentation of a string, while secondary come from decay of other particles. Since particles (top, say), may decay into partons, the distinction is not always unique. \iteme{= 22 :} gives a table of particle content in events. \iteme{= 23 :} stores particle content in events in \ttt{/LUJETS/}, using the following format: \\ \ttt{N =} number of different particle species found; \\ \ttt{K(I,1) =} 32; \\ \ttt{K(I,2) =} 99; \\ \ttt{K(I,3) =} particle KF code; \\ \ttt{K(I,5) =} total number of particles and antiparticles of this species; \\ \ttt{P(I,1) =} average number of primary particles per event; \\ \ttt{P(I,2) =} average number of secondary particles per event; \\ \ttt{P(I,3) =} average number of primary antiparticles per event; \\ \ttt{P(I,4) =} average number of secondary antiparticles per event; \\ \ttt{P(I,5) =} average total number of particles or antiparticles per event. \\ In addition, \ttt{MSTU(3)=1} and \\ \ttt{K(N+1,1) =} 32; \\ \ttt{K(N+1,2) =} 99; \\ \ttt{K(N+1,5) =} number of events analyzed; \\ \ttt{P(N+1,1) =} average primary multiplicity per event; \\ \ttt{P(N+1,2) =} average final multiplicity per event; \\ \ttt{P(N+1,3) =} average charged multiplicity per event. \iteme{= 30 :} statistics on factorial moments is reset. \iteme{= 31 :} analyzes the factorial moments of the multiplicity distribution in different bins of rapidity and azimuth. Which particles (or partons) are used in the analysis is determined by the \ttt{MSTU(41)} value. The selection between usage of true rapidity, pion rapidity or pseudorapidity is regulated by \ttt{MSTU(42)}. The $z$ axis is assumed to be event axis; if this is not desirable find an event axis e.g. with \ttt{LUSPHE} or \ttt{LUTHRU} and use \ttt{LUEDIT(31)}. Maximum (pion-, pseudo-) rapidity, which sets the limit for the rapidity plateau or the experimental acceptance, is given by \ttt{PARU(57)}. \iteme{= 32 :} prints a table of the first four factorial moments for various bins of pseudorapidity and azimuth. The moments are properly normalized so that they would be unity (up to statistical fluctuations) for uniform and uncorrelated particle production according to Poissonian statistics, but increasing for decreasing bin size in case of `intermittent' behaviour. The error on the average value is based on the actual statistical sample (i.e. does not use any assumptions on the distribution to relate errors to the average values of higher moments). Note that for small bin sizes, where the average multiplicity is small and the factorial moment therefore only very rarely is non-vanishing, moment values may fluctuate wildly and the errors given may be too low. \iteme{= 33 :} stores the factorial moments in \ttt{/LUJETS/}, using the format: \\ \ttt{N =} 30, with \ttt{I = }$i=1$--10 corresponding to results for slicing the rapidity range in $2^{i-1}$ bins, \ttt{I = }$i = 11$--20 to slicing the azimuth in $2^{i-11}$ bins, and \ttt{I = }$i = 21$--30 to slicing both rapidity and azimuth, each in $2^{i-21}$ bins; \\ \ttt{K(I,1) =} 32; \\ \ttt{K(I,2) =} 99; \\ \ttt{K(I,3) =} number of bins in rapidity; \\ \ttt{K(I,4) =} number of bins in azimuth; \\ \ttt{P(I,1) =} rapidity bin size; \\ \ttt{P(I,2) - P(I,5) =} $\langle F_2 \rangle$--$\langle F_5 \rangle$, i.e. mean of second, third, fourth and fifth factorial moment; \\ \ttt{V(I,1) =} azimuthal bin size; \\ \ttt{V(I,2) - V(I,5) =} statistical errors on $\langle F_2 \rangle$--$\langle F_5 \rangle$. \\ In addition, \ttt{MSTU(3) =} 1 and \\ \ttt{K(31,1) =} 32; \\ \ttt{K(31,2) =} 99; \\ \ttt{K(31,5) =} number of events analyzed. \iteme{= 40 :} statistics on energy--energy correlation is reset. \iteme{= 41 :} the energy--energy correlation $\mrm{EEC}$ of the current event is analyzed. Which particles (or partons) are used in the analysis is determined by the \ttt{MSTU(41)} value. Events are assumed given in their c.m. frame. The weight assigned to a pair $i$ and $j$ is $2 E_i E_j/E_{\mrm{vis}}^2$, where $E_{\mrm{vis}}$ is the sum of energies of all analyzed particles in the event. Energies are determined from the momenta of particles, with mass determined according to the \ttt{MSTU(42)} value. Statistics is accumulated for the relative angle $\theta_{ij}$, ranging between 0 and 180 degrees, subdivided into 50 bins. \iteme{= 42 :} prints a table of the energy--energy correlation $\mrm{EEC}$ and its asymmetry $\mrm{EECA}$, with errors. The definition of errors is not unique. In our approach each event is viewed as one observation, i.e. an $\mrm{EEC}$ and $\mrm{EECA}$ distribution is obtained by summing over all particle pairs of an event, and then the average and spread of this event-distribution is calculated in the standard fashion. The quoted error is therefore inversely proportional to the square root of the number of events. It could have been possible to view each single particle pair as one observation, which would have given somewhat lower errors, but then one would also be forced to do a complicated correction procedure to account for the pairs in an event not being uncorrelated (two hard jets separated by a given angle typically corresponds to several pairs at about that angle). Note, however, that in our approach the squared error on an $\mrm{EECA}$ bin is smaller than the sum of the squares of the errors on the corresponding $\mrm{EEC}$ bins (as it should be). Also note that it is not possible to combine the errors of two nearby bins by hand from the information given, since nearby bins are correlated (again a trivial consequence of the presence of jets). \iteme{= 43 :} stores the $\mrm{EEC}$ and $\mrm{EECA}$ in \ttt{/LUJETS/}, using the format: \\ \ttt{N =} 25; \\ \ttt{K(I,1) =} 32; \\ \ttt{K(I,2) =} 99; \\ \ttt{P(I,1) =} $\mrm{EEC}$ for angles between \ttt{I-1} and \ttt{I}, in units of $3.6^{\circ}$; \\ \ttt{P(I,2) =} $\mrm{EEC}$ for angles between \ttt{50-I} and \ttt{51-I}, in units of $3.6^{\circ}$; \\ \ttt{P(I,3) =} $\mrm{EECA}$ for angles between \ttt{I-1} and \ttt{I}, in units of $3.6^{\circ}$; \\ \ttt{P(I,4), P(I,5) :} lower and upper edge of angular range of bin \ttt{I}, expressed in radians; \\ \ttt{V(I,1) - V(I,3) :} errors on the $\mrm{EEC}$ and $\mrm{EECA}$ values stored in \ttt{P(I,1) - P(I,3)} (see \ttt{=42} for comments); \\ \ttt{V(I,4), V(I,5) :} lower and upper edge of angular range of bin \ttt{I}, expressed in degrees. \\ In addition, \ttt{ MSTU(3)=1} and \\ \ttt{K(26,1) =} 32; \\ \ttt{K(26,2) =} 99; \\ \ttt{K(26,5) =} number of events analyzed. \iteme{= 50 :} statistics on complete final states is reset. \iteme{= 51 :} analyzes the particle content of the final state of the current event record. During the course of the run, statistics is thus accumulated on how often different final states appear. Only final states with up to 8 particles are analyzed, and there is only reserved space for up to 200 different final states. Most high energy events have multiplicities far above 8, so the main use for this tool is to study the effective branching ratios obtained with a given decay model for e.g. charm or bottom hadrons. Then \ttt{LU1ENT} may be used to generate one decaying particle at a time, with a subsequent analysis by \ttt{LUTABU}. Depending on at what level this studied is to be carried out, some particle decays may be switched off, like $\pi^0$. \iteme{= 52 :} gives a list of the (at most 200) channels with up to 8 particles in the final state, with their relative branching ratio. The ordering is according to multiplicity, and within each multiplicity according to an ascending order of KF codes. The KF codes of the particles belonging to a given channel are given in descending order. \iteme{= 53 :} stores the final states and branching ratios found in \ttt{/LUJETS/}, using the format: \\ \ttt{N =} number of different explicit final states found (at most 200); \\ \ttt{K(I,1) =} 32; \\ \ttt{K(I,2) =} 99; \\ \ttt{K(I,5) =} multiplicity of given final state, a number between 1 and 8; \\ \ttt{P(I,1) - P(I,5), V(I,1) - V(I,3) :} the KF codes of the up to 8 particles of the given final state, converted to real numbers, with trailing zeroes for positions not used; \\ \ttt{V(I,5) :} effective branching ratio for the given final state. \\ In addition, \ttt{MSTU(3)=1} and \\ \ttt{K(N+1,1) =} 32; \\ \ttt{K(N+1,2) =} 99; \\ \ttt{K(N+1,5) =} number of events analyzed; \\ \ttt{V(N+1,5) =} summed branching ratio for finals states not given above, either because they contained more than 8 particles or because all 200 channels have been used up. \end{subentry} \end{entry} \drawbox{COMMON/LUDAT1/MSTU(200),PARU(200),MSTJ(200),PARJ(200)} \begin{entry} \itemc{Purpose:} to give access to a number of status codes and parameters which regulate the performance of {\tsc{Jetset}}. Most parameters are described in section \ref{ss:JETswitch}; here only those related to the event-analysis routines are described. \boxsep \iteme{MSTU(41) :}\label{p:MSTU41} (D=2) partons/particles used in the event-analysis routines \ttt{LUSPHE}, \ttt{LUTHRU}, \ttt{LUCLUS}, \ttt{LUCELL}, \ttt{LUJMAS}, \ttt{LUFOWO} and \ttt{LUTABU} (\ttt{LUTABU(11)} excepted). \begin{subentry} \iteme{= 1 :} all partons/particles that have not fragmented/decayed. \iteme{= 2 :} ditto, with the exception of neutrinos and unknown particles. \iteme{= 3 :} only charged, stable particles, plus any partons still not fragmented. \end{subentry} \iteme{MSTU(42) :} (D=2) assumed particle masses, used in calculating energies $E^2 = \mbf{p}^2 + m^2$, as subsequently used in \ttt{LUCLUS}, \ttt{LUJMAS} and \ttt{LUTABU} (in the latter also for pseudorapidity, pion rapidity or true rapidity selection). \begin{subentry} \iteme{= 0 :} all particles are assumed massless. \iteme{= 1 :} all particles, except the photon, are assumed to have the charged pion mass. \iteme{= 2 :} the true masses are used. \end{subentry} \iteme{MSTU(43) :} (D=1) storing of event-analysis information (mainly jet axes), in \ttt{LUSPHE}, \ttt{LUTHRU}, \ttt{LUCLUS} and \ttt{LUCELL}. \begin{subentry} \iteme{= 1 :} stored after the event proper, in positions \ttt{N+1} through \ttt{N+MSTU(3)}. If several of the routines are used in succession, all but the latest information is overwritten. \iteme{= 2 :} stored with the event proper, i.e. at the end of the event listing, with \ttt{N} updated accordingly. If several of the routines are used in succession, all the axes determined are available. \end{subentry} \iteme{MSTU(44) :} (D=4) is the number of the fastest (i.e. with largest momentum) particles used to construct the (at most) 10 most promising starting configurations for the thrust axis determination. \iteme{MSTU(45) :} (D=2) is the number of different starting configurations above, which have to converge to the same (best) value before this is accepted as the correct thrust axis. \iteme{MSTU(46) :} (D=1) distance measure used for the joining of clusters in \ttt{LUCLUS}. \begin{subentry} \iteme{= 1 :} $d_{ij}$, i.e. approximately relative transverse momentum. Anytime two clusters have been joined, particles are reassigned to the cluster they now are closest to. The distance cut-off $d_{\mrm{join}}$ is stored in \ttt{PARU(44)}. \iteme{= 2 :} distance measure as in \ttt{=1}, but particles are never reassigned to new jets. \iteme{= 3 :} JADE distance measyre $y_{ij}$, but with dimensions to correspond approximately to total invariant mass. Particles may never be reassigned between clusters. The distance cut-off $m_{\mathrm{min}}$ is stored in \ttt{PARU(44)}. \iteme{= 4 :} as \ttt{=3}, but a scaled JADE distance $y_{ij}$ is used instead of $m_{ij}$. The distance cut-off $y_{\mathrm{min}}$ is stored in \ttt{PARU(45)}. \end{subentry} \iteme{MSTU(47) :} (D=1) the minimum number of clusters to be reconstructed by \ttt{LUCLUS}. \iteme{MSTU(48) :} (D=0) mode of operation of the \ttt{LUCLUS} routine. \begin{subentry} \iteme{= 0 :} the cluster search is started from scratch. \iteme{= 1 :} the clusters obtained in a previous cluster search on the same event (with \ttt{MSTU(48)=0}) are to be taken as the starting point for subsequent cluster joining. For this call to have any effect, the joining scale in \ttt{PARU(44)} or \ttt{PARU(45)} must have been changed. If the event record has been modified after the last \ttt{LUCLUS} call, or if any other cluster search parameter setting has been changed, the subsequent result is unpredictable. \end{subentry} \iteme{MSTU(51) :} (D=25) number of pseudorapidity bins that the range between \ttt{-PARU(51)} and \ttt{+PARU(51)} is divided into to define cell size for \ttt{LUCELL}. \iteme{MSTU(52) :} (D=24) number of azimuthal bins, used to define the cell size for \ttt{LUCELL}. \iteme{MSTU(53) :} (D=0) smearing of correct energy, imposed cell-by-cell in \ttt{LUCELL}, to simulate calorimeter resolution effects. \begin{subentry} \iteme{= 0 :} no smearing. \iteme{= 1 :} the transverse energy in a cell, $E_{\perp}$, is smeared according to a Gaussian distribution with standard deviation \ttt{PARU(55)}$\times \sqrt{E_{\perp}}$, where $E_{\perp}$ is given in GeV. The Gaussian is cut off so that $0 < E_{\perp \mrm{smeared}} < $\ttt{PARU(56)}$\times E_{\perp \mrm{true}}$. \iteme{= 2 :} as \ttt{=1}, but it is the energy $E$ rather than the transverse energy $E_{\perp}$ that is smeared. \end{subentry} \iteme{MSTU(54) :} (D=1) form for presentation of information about reconstructed clusters in \ttt{LUCELL}, as stored in \ttt{LUJETS} according to the \ttt{MSTU(43)} value. \begin{subentry} \iteme{= 1 :} the \ttt{P} vector in each line contains $\eta$ and $\varphi$ for the geometric origin of the jet, $\eta$ and $\varphi$ for the weighted center of the jet, and jet $E_{\perp}$, respectively. \iteme{= 2 :} the \ttt{P} vector in each line contains a massless four-vector giving the direction of the jet, obtained as \\ $(p_x,p_y,p_z,E,m) = E_{\perp} (\cos\varphi,\sin\varphi,\sinh\eta,\cosh\eta,0)$, \\ where $\eta$ and $\varphi$ give the weighted center of a jet and $E_{\perp}$ its transverse energy. \iteme{= 3 :} the \ttt{P} vector in each line contains a massive four-vector, obtained by adding the massless four-vectors of all cells that form part of the jet, and calculating the jet mass from $m^2 = E^2-p_x^2-p_y^2-p_z^2$. For each cell, the total $E_{\perp}$ is summed up, and then translated into a massless four-vector assuming that all the $E_{\perp}$ was deposited in the center of the cell. \end{subentry} \iteme{MSTU(61) :} (I) first entry for storage of event-analysis information in last event analyzed with \ttt{LUSPHE}, \ttt{LUTHRU}, \ttt{LUCLUS} or \ttt{LUCELL}. \iteme{MSTU(62) :} (R) number of particles/partons used in the last event analysis with \ttt{LUSPHE}, \ttt{LUTHRU}, \ttt{LUCLUS}, \ttt{LUCELL}, \ttt{LUJMAS}, \ttt{LUFOWO} or \ttt{LUTABU}. \iteme{MSTU(63) :} (R) in a \ttt{LUCLUS} call, the number of preclusters constructed in order to speed up analysis (should be equal to \ttt{MSTU(62)} if \ttt{PARU(43)=0.}). In a \ttt{LUCELL} call, the number of cells hit. \iteme{MSTU(161), MSTU(162) :}\label{p:MSTU161} hard flavours involved in current event, as used in an analysis with \ttt{LUTABU(11)}. Either or both may be set 0, to indicate the presence of one or none hard flavours in event. Is normally set by high-level routines, like \ttt{LUEEVT} or \ttt{PYEVNT}, but can also be set by you. \boxsep \iteme{PARU(41) :}\label{p:PARU41} (D=2.) power of momentum-dependence in \ttt{LUSPHE}, default corresponds to sphericity, \ttt{=1.} to linear event measures. \iteme{PARU(42) :} (D=1.) power of momentum-dependence in \ttt{LUTHRU}, default corresponds to thrust. \iteme{PARU(43) :} (D=0.25 GeV) maximum distance $d_{\mrm{init}}$ allowed in \ttt{LUCLUS} when forming starting clusters used to speed up reconstruction. The meaning of the parameter is in $p_{\perp}$ for \ttt{MSTU(46)}$\leq 2$ and in $m$ for \ttt{MSTU(46)}$\geq 3$. If \ttt{=0.}, no preclustering is obtained. If chosen too large, more joining may be generated at this stage than is desirable. The main application is at high energies, where some speedup is imperative, and the small details are not so important anyway. \iteme{PARU(44) :} (D=2.5 GeV) maximum distance $d_{\mrm{join}}$, below which it is allowed to join two clusters into one in \ttt{LUCLUS}. Is used for \ttt{MSTU(46)}$\leq 3$, i.e. both for $p_{\perp}$ and mass distance measure. \iteme{PARU(45) :} (D=0.05) maximum distance $y_{\mrm{join}} = m^2/E_{\mrm{vis}}^2$, below which it is allowed to join two clusters into one in \ttt{LUCLUS} for \ttt{MSTU(46)=4}. \iteme{PARU(48) :} (D=0.0001) convergence criterion for thrust (in \ttt{LUTHRU}) or generalized thrust (in \ttt{LUCLUS}), or relative change of $m_{\mrm{H}}^2 + m_{\mrm{L}}^2$ (in \ttt{LUJMAS}), i.e. when the value changes by less than this amount between two iterations the process is stopped. \iteme{PARU(51) :} (D=2.5) defines maximum absolute pseudorapidity used for detector assumed in \ttt{LUCELL}. \iteme{PARU(52) :} (D=1.5 GeV) gives minimum $E_{\perp}$ for a cell to be considered as a potential jet initiator by \ttt{LUCELL}. \iteme{PARU(53) :} (D=7.0 GeV) gives minimum summed $E_{\perp}$ for a collection of cells to be accepted as a jet. \iteme{PARU(54) :} (D=1.) gives the maximum distance in $R = \sqrt{(\Delta\eta)^2 + (\Delta\varphi)^2}$ from cell initiator when grouping cells to check whether they qualify as a jet. \iteme{PARU(55) :} (D=0.5) when smearing the transverse energy (or energy, see \ttt{MSTU(53)}) in \ttt{LUCELL}, the calorimeter cell resolution is taken to be \ttt{PARU(55)}$\times \sqrt{E_{\perp}}$ (or \ttt{PARU(55)}$\times \sqrt{E}$) for $E_{\perp}$ (or $E$) in GeV. \iteme{PARU(56) :} (D=2.) maximum factor of upward fluctuation in transverse energy or energy in a given cell when calorimeter resolution is included in \ttt{LUCELL} (see \ttt{MSTU(53)}). \iteme{PARU(57) :} (D=3.2) maximum rapidity (or pseudorapidity or pion rapidity, depending on \ttt{MSTU(42)}) used in the factorial moments analysis in \ttt{LUTABU}. \iteme{PARU(58) :} (D=0. GeV) in a \ttt{LUCELL} call, cells with a transverse energy $E_{\perp}$ below \ttt{PARP(58)} are removed from further consideration. This may be used to represent a threshold in an actual calorimeter, or may be chosen just to speed up the algorithm in a high-multiplicity environment.. \iteme{PARU(61) :} (I) invariant mass $W$ of a system analyzed with \ttt{LUCLUS} or \ttt{LUJMAS}, with energies calculated according to the \ttt{MSTU(42)} value. \iteme{PARU(62) :} (R) the generalized thrust obtained after a successful \ttt{LUCLUS} call, i.e. ratio of summed cluster momenta and summed particle momenta. \iteme{PARU(63) :} (R) the minimum distance $d$ between two clusters in the final cluster configuration after a successful \ttt{LUCLUS} call; is 0 if only one cluster left. \end{entry} \clearpage \section{Summary and Outlook} A complete description of the {\tsc{Pythia/Jetset}} programs would have to cover four aspects: \begin{Enumerate} \item the basic philosophy and principles underlying the programs; \item the detailed physics scenarios implemented, with all the necessary compromises and approximations; \item the structure of the implementation, including program flow, internal variable names and programming tricks; and \item the manual, which describes how to use the programs. \end{Enumerate} Of these aspects, the first has been dealt with in reasonable detail. The second is unevenly covered: in depth for aspects which are not discussed anywhere else, more summarily for areas where separate up-to-date papers already exist. The third is not included at all, but `left as an exercise' for the reader, to figure out from the code itself. The fourth, finally, should be largely covered, although many further comments could have been made, in particular about the interplay between different parts of the programs. Still, in the end, no manual, however complete, can substitute for `hands on' experience. The {\tsc{Pythia/Jetset}} programs are continuously being developed. We are aware of many shortcomings, which hopefully will be addressed in the future, such as: \begin{Itemize} \item polarization effects should be included in more places, in particular for $\tau$ production and decay; \item the photoproduction and $\gamma\gamma$-physics scenarios should be expanded; \item many processes of interest are missing; and \item mass relations and couplings need to be included beyond the Born level in the MSSM two Higgs doublet scenario. \end{Itemize} This list could have been made much longer (I almost certainly missed your top priority). One other aspect would be to provide more and longer examples of working main programs for a number of standard applications. Apart from these physics aspects, one may also worry about the programming ones. For instance, for historical reasons, single precision real is used almost everywhere. With the push to higher energies, this is becoming more and more of a problem, so it would be logical to move to double precision throughout. One should also note that the {\tsc{Jetset}} and {\tsc{Pythia}} programs these days are becoming so intertwined, that it would make sense to join them into one single program. This would e.g. mean that the current $\e^+\e^-$ generation routines of {\tsc{Jetset}} are made part of the generic {\tsc{Pythia}} process generation machinery --- this is particular affects the matrix-element options, since $\e^+\e^-$ events with parton showers already are available in {\tsc{Pythia}}. A joint product would likely adopt the name {\tsc{Pythia}}: although {\tsc{Jetset}} is the older of the two programs, it has a less well developed identity of its own. (It is also often referred to as `Lund', which today is more confusing than it was in the early days.) In the process of joining the programs, one would probably also remove a number of options that are no longer used. Another possible change on longer time scales would be an introduction of Fortran 90 programming elements. In particular, derived data types could be used to define the event record as a one-dimensional array, where each element represents a particle, with integer and real components to give flavour, history, momentum and production vertex. No timetable is set up for future changes. After all, this is not a professionally maintained software product, but part of a one-man physics research project. Very often, developments of the programs have come about as a direct response to the evolution of the physics stage, i.e. experimental results and studies for future accelerators. Hopefully, the program will keep on evolving in step with the new challenges opening up. \clearpage
\section{Introduction} Photon-proton processes are traditionally classified according to the virtuality ($Q^2$) of the photon. For quasi-real photoproduction interactions $Q^2$ is close to zero, and correspondingly the photon is nearly on mass shell. Interactions which involve a $Q^2$ larger than a few GeV$^2$ are usually termed deep-inelastic scattering processes (DIS). This distinction results mainly from the different theoretical descriptions adopted for these processes. Photoproduction has turned out to be very similar to hadron-hadron collisions and is described in a VMD-like (Vector Meson Dominance) picture~\cite{sakurai}, where the photon is assumed to fluctuate into a vector meson before interacting with the proton. Additional diagrams such as the direct coupling of the photon to quarks in the proton and a pointlike contribution where the photon splits into a $q\overline{q}$ pair are needed to accommodate the large tail observed in the transverse momentum distribution of produced particles~\cite{H1:incl}, but the majority of photoproduction collisions show features of soft low-$p_T$ processes as seen in hadronic collisions. For DIS on the other hand, the observation of scaling of the structure function in early experiments suggested that this interaction could, in the infinite momentum frame of the proton, be interpreted as a hard scattering process in which a point-like virtual photon ``probes'' the structure of the hadronic target. The subsequently measured scaling violations are well described by perturbative QCD. However, recent DIS measurements at the high energy $ep$ collider HERA and high precision measurements from fixed target leptoproduction experiments have shown that the data in the low Bjorken-$x$ region reveal properties from soft interactions as well. Here $x = Q^2/2P\cdot q$ where $P$ and $q$ are the four momenta of the incoming proton and photon, respectively. Elastic and other diffractive hadronic final states have been produced~\cite{h1rapgap,disdif} and shadowing on nuclear targets has been observed~\cite{shadowing}, phenomena which are well known from hadron-hadron and real photon collisions~\cite{gvmd}. These observations suggest that the different treatment of low and high $Q^2$ interactions is somewhat artificial, and it is therefore worthwhile testing a prescription which provides a natural transition between these two classes. A qualitative approach for a smooth transition of the high energy $\gamma^*p$~\footnote{In this paper the generic symbol $\gamma^*$ is used to denote a colliding photon irrespective of the virtuality.} interactions over a large range of $Q^2$ results from viewing the photon-proton collision in the proton rest frame and the hypothesis that the photon can develop as a hadronic object before interacting with the nuclear target. The time in which a real photon can fluctuate into e.g. a $\rho$ meson is given by the Heisenberg uncertainty principle and amounts to $\tau \approx 2\nu/M^2_{\rho}$ where $\nu$ is the photon energy in the proton rest frame and $M_{\rho}$ the mass of the $\rho$ meson. The lifetime of a high energy real photon to fluctuate into a hadronic state is much longer than the time of the strong interaction itself, and therefore this picture is generally applied to describe photoproduction interactions. The same argument can be used for DIS interactions at low $x$. In the proton rest frame the time in which the virtual photon can fluctuate and stay in a hadronic state, e.g. a quark-antiquark pair, is given by $\tau \approx 1/(xM_p)$ ~\cite{tioffe,brodsky}, where $M_p$ is the mass of the nucleon target. Thus, for small $x$, the virtual photon can convert into a quark-antiquark pair and cover a distance which is long compared to the interaction length. For an $x$ in the range of $10^{-2}-10^{-4}$ this distance is in the range of 10 to 1000 fm, much larger than the typical radius of the target. In the HERA kinematic range such $x$ values can be reached for $Q^2 \sim 10 $ GeV$^2$. Therefore, at low $x$, a virtual photon can stay in a hadronic state for a long time and interact with the target strongly, leading to a final state similar to the one in a hadron-hadron collision. This picture of DIS in the frame where the incoming photon fluctuates into a hadronic system before interacting with the proton has already been advocated in various papers~\cite{nikolaev,alligned,kwiecinski}. Note that this is not in contradiction with the traditional treatment of DIS as a point-like process, where a virtual photon ``probes" the hadronic structure of the nuclear target. In fact both pictures are taken to be complementary, as discussed in~\cite{brodsky}. The HERA collider provides a unique opportunity to study the final state of both photoproduction and DIS interactions at high energy and small Bjorken-$x$: 26.7 GeV electrons collide with 820 GeV protons, yielding a centre of mass energy $\sqrt{s}$ of 296 GeV. This allows for a study of DIS interactions for $x$ values down to $\sim 10^{-4}$~\cite{h1f2}, and photoproduction collisions at a centre of mass energy of $\sim 200$ GeV~\cite{H1:incl}. The data recorded with the H1 experiment are used to study the transverse energy behaviour in photoproduction and DIS interactions, and the analogy with hadronic collisions is checked. Within this analogy $\mbox{$\gamma^{*}p$}$ events can be sub-divided in rapidity-space into three regions, which differ in the mechanism by which the hadronic final state is produced~\cite{bj73}: the proton fragmentation region, the photon fragmentation region and a hadron plateau spanning the rapidity interval between the two. The height of the hadronic plateau depends logarithmically on the centre of mass energy of the $\mbox{$\gamma^{*}p$}$ collision. The region of the hadronic plateau is expected to be independent of the nature of the two incoming ``hadrons", as was verified in hadron-hadron collisions~\cite{hadver}. Particle production in the proton fragmentation region is also expected to be very similar compared to hadron-proton collisions, but the limited experimental acceptance in this region does not allow this to be verified. The ``hadronic" nature of the photon is assumed to change with increasing $Q^2$, thus the photon fragmentation region is expected to change with $Q^2$. Also studied is the fraction of diffractive events in DIS and photoproduction. In hadronic collisions approximate factorization of the cross section for high mass diffractive dissociation has been observed, in accordance with Regge theory predictions~\cite{triple}: the ratio of the hadron diffractive dissociation to the total cross section is found to be approximately independent of the type of dissociating hadron and is thus the same for pions, kaons, protons and real photons~\cite{triple2}. Applying this factorization property to real and virtual photons leads to the expectation that the ratio of the photon diffractive dissociation cross section to the total cross section is independent of the $Q^2$ of the photon, in the high dissociative mass region. \section{Detector Description} A detailed description of the H1 apparatus can be found elsewhere~\cite{h1}. In the following the components of the detector relevant for this analysis are briefly described. The hadronic energy flow and the scattered electrons in DIS processes are measured with a liquid argon~(LAr) calorimeter and a backward electromagnetic lead-scintillator calorimeter (BEMC). The LAr calorimeter~\cite{larc} covers the polar angular range $4^\circ < \theta < 153^\circ$ with full azimuthal coverage, where $\theta$ is defined with respect to the proton beam direction ($+z$ axis). It consists of an electromagnetic section with lead absorbers and a hadronic section with steel absorbers. Both sections are highly segmented in the transverse and longitudinal direction with about 44000 cells in total. The total depth of both sections varies between 4.5 and 8 interaction lengths. Test beam measurements of the LAr~calorimeter modules show an energy resolution of $\sigma_{E}/E\approx 0.12/\sqrt{E\;[\rm GeV]} \oplus 0.01$ for electrons ~\cite{elcern} and $\sigma_{E}/E\approx 0.50/\sqrt{E\;[\rm GeV]} \oplus 0.02$ for charged pions~\cite{h1calo3}. The uncertainty in the absolute energy scale for electrons is $3\%$. The absolute scale of the hadronic energy is presently known to $5\%$, as determined from studies of the transverse momentum balance in DIS events. The BEMC (depth of 22.5 radiation lengths or 1 interaction length) covers the backward region of the detector, $151^\circ < \theta < 177^\circ$. A major task of the BEMC is to trigger on and measure scattered electrons in DIS processes with $Q^2$ values ranging from 5 to 100 GeV$^{2}$. The BEMC energy scale for electrons is known to an accuracy of $1.7\%$~\cite{bemc}. Its resolution is given by $\sigma_{E}/E = 0.10/\sqrt{E\;[\rm GeV]} \oplus 0.42/E[\rm GeV] \oplus 0.03$ \cite{f2pap}. The calorimeters are surrounded by a superconducting solenoid providing a uniform magnetic field of $1.15$ T parallel to the beam axis in the tracking region. Charged particle tracks are measured in a central drift chamber and the forward tracking system, covering the polar angular range $ 7^\circ < \theta < 165^\circ$. The central chamber is interleaved with an inner and an outer double layer of multi-wire proportional chambers (MWPC), which were used for the trigger to select events with charged tracks pointing to the interaction region. A backward proportional chamber (BPC) in front of the BEMC with an angular acceptance of $155.5^\circ < \theta < 174.5^\circ$ serves to support electron identification and to precisely measure electron direction. Using information from the BPC, the BEMC and the reconstructed event vertex the polar angle of the scattered electron is known to better than 2 mrad. A small angle detector (electron tagger), which is part of the luminosity system, is a TlCl/TlBr crystal calorimeter with an energy resolution $\sigma_{E}/E = 0.1/\sqrt{E\;[\rm GeV]}$. It is located at $z$ = --33 m and accepts electrons from photoproduction processes with an energy fraction between 0.2 and 0.8 with respect to the beam energy and scattering angles $\theta'< 5$~mrad ($\theta' = \pi-\theta$). \section{Event Selection and Correction for Detector Effects} The data used in this analysis were collected in 1993 and correspond to an integrated luminosity of about $0.3$ pb$^{-1}$. The kinematic variables $Q^2, x$ and $y$ of the $ep$ collision are determined from the scattered electron: $ Q^2 = 4\,E_e \, E'_e\cos^2(\theta_{e}/2)$ and $ y = 1-(E'_e/E_e)\cdot \sin^2(\theta_{e}/2)$, where $E_e$ is the energy of the incident electron and $E_e'$ and $\theta_e$ are the energy and the polar angle of the scattered electron respectively. The variable $y$ represents the fraction of the energy of the electron transferred to the proton, in the proton rest frame. The scaling variable $x$ is then derived via $x=Q^2/(ys)$, and the total hadronic invariant mass is given by $W^2=sy-Q^2$. \begin{table}[thb] \centering \begin{tabular}{|c|c|c|} \hline photoproduction & low $Q^2$ DIS & high $Q^2$ DIS \\ sample & sample & sample \\ \hline $Q^2<10^{-2}$~GeV$^2$ & $5<Q^2<100$~GeV$^2$ & $Q^2>100$~GeV$^2$ \\ \hline $\theta_e-\pi < 5$~mrad & $157^\circ<\theta_e < 173^\circ$ & $10^\circ < \theta_e < 148^\circ$ \\ \hline $0.3 < y < 0.5$ & $E'_e > 12$~GeV & $0.05< y < 0.7$ \\ & $y > 0.05$ & \\ \hline \end{tabular} \caption[~] {\small Accepted kinematic regions for the three sub-samples used in this analysis. } \label{seltable} \end{table} The data are classified into three event sub-samples depending on the $Q^2$ range. A different detector component of H1 is used to detect the scattered electron in each of these ranges. In the studies below these subsamples get further sub-divided into samples with different $y$ or $Q^2$ values. \begin{itemize} \item The photoproduction sub-sample ($Q^2<10^{-2}$~GeV$^2$) consists of events where the scattered electron is detected in the small angle electron detector. To avoid regions of low acceptance and to facilitate the data correction procedure, the kinematic range was further restricted to $0.3<y<0.5$. The photoproduction events are triggered by a coincidence of an energy deposit in the small angle electron detector and at least one track pointing to the vertex region. The track condition is derived from the cylindrical MWPC and requires a $p_T\;\raisebox{-0.5mm}{$\stackrel{>}{\scriptstyle{\sim}}$}\;200$ MeV/c. \item The low $Q^2$ DIS sub-sample ($5<Q^2<100$~GeV$^2$) consists of events where the scattered electron is detected in the BEMC. The events are triggered by requiring a cluster of more than 4~GeV in the BEMC. The most energetic BEMC cluster is taken to be the scattered electron. \item For the high $Q^2$ DIS sub-sample ($Q^2>100$~GeV$^2$) the scattered electron is detected in the LAr calorimeter. The events are triggered by requiring a cluster of more than 5~GeV in the electromagnetic part of the LAr calorimeter, and no associated hadronic energy. The electromagnetic cluster in the LAr calorimeter with the highest transverse energy is considered to be the scattered electron. \end{itemize} Further details on the scattered electron identification procedure in the BEMC and LAr calorimeters and in the small angle electron detector can be found in~\cite{h1f2,h1alphas,h1gamp}, respectively. The selected kinematic regions for each sub-sample, as given in Table~\ref{seltable}, are chosen to ensure a large acceptance, high trigger efficiency and, for the DIS sub-samples, a small photoproduction background (less than 3\%). For all three sub-samples the $z$ position of the event vertex reconstructed from charged tracks was required to be within $\pm30$~cm of the nominal interaction point. A minimum $y$ cut, $y > 0.05$ was imposed for the DIS sub-samples. Events suffering from QED radiation or from a badly reconstructed electron are strongly reduced by requiring that they also fulfill this cut if $y$ is calculated from the measured hadrons. The final event samples contain 82850 photoproduction events, 15324 low $Q^2$ DIS events with $5 < Q^2 < 100$~GeV$^2$ ($10^{-4} < x < 10^{-2}$), 692 high $Q^2$ DIS events with $Q^2 > 100$~GeV$^2$ ($10^{-3} < x < 10^{-1}$). The data are corrected for detector effects using samples of Monte Carlo generated events which were fully simulated in the H1 detector. The PHOJET~\cite{phojet} generator for photoproduction and the CDM~\cite{ariadne} (Colour Dipole Model) generator for DIS processes were used. The program PHOJET generates $\gamma p$ interactions, treating the photon as a hadron-like object. The model used to simulate the hadronic final states is similar to that used in the Monte Carlo program DTUJET~\cite{dtujet} which simulates particle production in $pp$ and $\bar{p}p$ collisions up to very high energies. The CDM Monte Carlo program generates DIS events, and uses the colour dipole model for QCD radiation in the hadronic final state. Here the final state is assumed to be a chain of independently radiating dipoles formed by emitted gluons~\cite{dipole}. Since all radiation is assumed to originate from the dipole formed by the struck quark and the remnant, photon-gluon fusion events have to be added and are taken from the QCD matrix elements~\cite{lepto}. Version 4.03 of the ARIADNE program was used for the CDM studies in this paper. To estimate systematic uncertainties of the correction procedure for DIS events another model for the hadronic final state was used as well: the MEPS (Matrix Elements plus Parton Showers) model~\cite{lepto}. This model incorporates QCD matrix elements up to first order, with additional soft emissions generated by adding leading log parton showers. Divergences of the matrix element are avoided by imposing a lower limit on the parton-parton invariant masses, which was parametrized as a function of $W$ such that it is always 2 GeV above the region in phase space where the order $\alpha_s$ contributions would exceed the total cross section. More details on this implementation are given in~\cite{h1gljet}. \section{Results} For comparisons of event properties at different $Q^2$ values, the $\gamma^*p$ centre of mass system (CMS) is chosen as frame of reference. The orientation of the CMS is such that the direction of the proton defines the positive $z'$ axis. Transverse quantities are defined with respect to the proton direction in this frame. The flow of transverse energy, $E_T$, as a function of pseudorapidity $\eta^*=-\ln(\tan \frac{\theta^*}{2})$ in the CMS is shown in Fig.~\ref{ETA}. Here $\theta^*$ is the polar angle of a particle in the CMS frame with respect to the proton direction. The transverse energy is calculated from the energy deposits in the calorimeter cells. For this study the DIS data are restricted to the same kinematic region $0.3<y<0.5$ as the photoproduction events. The $y$ range corresponds to a mean $\mbox{$\gamma^{*}p$}$ collision energy of $\sqrt{s_{\gamma p}} = W \approx 185$~GeV. The corrections which have been applied to correct for detector effects never exceed 30$\%$. \begin{figure}[htb] \centering \Large \boldmath \begin{picture}(170,110)(-8,0) \put( 26,98){\small $\average{Q^2}\sim 0 $ GeV$^2$} \put( 26,92){\small $\average{Q^2}\sim 11 $ GeV$^2$} \put( 26,86){\small $\average{Q^2}\sim 38 $ GeV$^2$} \put( 26,80){\small $\average{Q^2}\sim 520 $ GeV$^2$} \put( 115,5){$\eta^* $} \put( 5,+42){\begin{sideways} $ 1/N \cdot dE_T/d\eta^*~~[$GeV$]$ \end{sideways}} \epsfig {file=fig1.ps,height=130mm,width=130mm,angle=90} \end{picture} \normalsize \unboldmath \caption[~] {\small The flow of transverse energy $E_T$ in the hadronic CMS as a function of pseudorapidity $\eta^*$ normalized to the number of events $N$. Photoproduction data (open circles) are compared with DIS data (full symbols: circles -- $\average{Q^2}\approx 11$~GeV$^2$, triangles -- $\average{Q^2}\approx 38$~GeV$^2$, squares -- $\average{Q^2}\approx 520$~GeV$^2$) in the same $y$ range $0.3 < y < 0.5$ ($W\approx 185$ GeV). } \label{ETA} \end{figure} \begin{figure}[htb] \centering \Large \boldmath \begin{picture}(170,110)(7,0) \put( 7,+45){\begin{sideways} $1/N\cdot dE_T/d\eta^* ~[$GeV$]$ \end{sideways}} \put( 38,89){\small $-3<\eta^*<-2$} \put( 38,96){\small $-0.5<\eta^*<0.5$} \put( 33,82){\small -------- CDM} \put( 33,76){\small -- -- -- MEPS} \put(125,5){$Q^2~~[$GeV$^2]$ } \put( 36,18){\large $\wr$} \put( 37,18){\large $\wr$} \put( 36,107){\large $\wr$} \put( 37,107){\large $\wr$} \put( 26,15){\small \bf 0} \epsfig {file=fig2.ps,height=130mm,width=170mm,angle=90} \end{picture} \normalsize \unboldmath \caption[~] {\small The transverse energy per unit of pseudorapidity in the CMS central region ($-0.5<\eta^*<0.5$, full circles) and in the photon fragmentation region ($-3<\eta^*<-2$, open circles) as a function of $Q^2$ for $0.3 < y < 0.5$. For comparison, the CDM (full line) and MEPS (dashed line) models are shown. The lower curves correspond to the CMS central region and the upper ones to the photon fragmentation region. } \label{Q2} \end{figure} \begin{figure}[htb] \centering \Large \boldmath \begin{picture}(170,110)(7,0) \put( 40,97){\large H1, $Q^2\approx~8~$GeV$^2$} \put( 40,92){\large H1, $Q^2\approx~14~$GeV$^2$} \put( 40,87){\large H1, $Q^2\approx~30~$GeV$^2$} \put( 40,82){\large H1, photoproduction} \put( 40,77){\large UA1} \put( 40,72){\large AFS} \put( 40,67){\large NA22} \put( 7,20){\begin{sideways} $1/N\cdot dE_T/d\eta^*$ at $\eta^*=0~[$GeV$]$ \end{sideways}} \put(125,5){$ W ~~[$GeV$]$ } \epsfig {file=fig3.ps,height=130mm,width=170mm,angle=90} \end{picture} \normalsize \unboldmath \caption[~] {\small The transverse energy per unit of pseudorapidity in the CMS central region as a function of the hadronic CMS energy. The DIS data are compared with photoproduction data and with data from hadron-hadron collisions ($p\overline{p}$ for UA1; $pp$ for NA22 and AFS). Systematic point-to-point errors of $6\%$ and an overall scale error of $9\%$ for the H1 data are not shown, neither are the global scale uncertainties for the other experiments. } \label{W} \end{figure} \begin{figure}[htb] \centering \Large \boldmath \begin{picture}(170,110)(7,0) \put( 25,100){\Large $a)$} \put(105,100){\Large $b)$} \put( 50,5){$E_T~~[$GeV$]$} \put( 5,+30){\begin{sideways} $ 1/N \cdot dN/dE_T~~[$GeV$^{-1}]$ \end{sideways}} \put( 140,5){$\eta^*_{max} $} \put(85,+45){\begin{sideways} $ 1/N \cdot dN/d\eta^*_{max}$ \end{sideways}} \epsfig {file=fig4.ps,height=130mm,width=170mm,angle=90} \end{picture} \normalsize \unboldmath \caption[~] {\small $(a)$ distribution of the uncorrected transverse energy per unit of pseudorapidity in the central region $(-0.5<\eta^*<0.5)~$, $(b)$ $\eta^*_{max}$ distribution normalized to the total number of events. Open circles - photoproduction data, full circles - DIS data with~$10<Q^2<100$~GeV$^2$ and $0.3<y<0.5$.} \label{ET-MAX} \end{figure} Only statistical errors are shown in Fig.~\ref{ETA}. For the comparison of the $E_T$ spectra at the four different $Q^2$ values it is important to identify the relative systematic uncertainty, so the systematic errors on each ``$Q^2$ point" at a given $\eta^*$ value are considered in two parts. Firstly there is a point-to-point systematic error, which has a different effect on each of the four measurements, accounting for kinematically dependent systematic effects. Secondly there is an overall scale error which affects all four points at a given $\eta^*$ value in the same way. The main source of the point-to-point systematic error is the model dependence of the correction for detector effects. This dependence was investigated with the CDM and MEPS models and leads to a point-to-point error of $6\%$\footnote{In the central region the correction for photoproduction events determined with PHOJET is close to the one for DIS events determined with MEPS at $<Q^2>\sim 11$ GeV$^2$}. Additionally, the model dependence contributes $6\%$ to the overall scale error at each $\eta^*$ value. Further contributions to the overall scale error arise from the LAr and BEMC calorimeter calibration ($5\%$ and $20\%$ respectively), and from details of the analysis method, such as the calorimeter noise treatment, the clustering scheme for the calorimeter cells, and the accuracy of the simulation of the calorimeter response, affecting the results by $5\%$ in the region close to the proton direction ($\eta^* > -1$). All these contributions give an overall scale error varying from $8\%$ in the photon fragmentation region ($-3.5 < \eta^* < -1$) to $9\%$ in the central region ($\eta^* > -1$), and up to $20\%$ in the region $\eta^* < -3.5$, which does not affect the results of the present analysis. This discussion on the systematic errors is also valid for data shown below. Fig.\ref{ETA} demonstrates that the transverse energy flow exhibits a strong increase in the photon fragmentation region ($\eta^* < -1$), from $Q^2=0$ (photoproduction) to $Q^2 \approx 500$~GeV$^2$ (DIS), while in the central region the level of $E_T$ remains almost the same. This can be taken as evidence that the influence of the $Q^2$ of the photon on the $E_T$ flow diffuses away quite quickly towards the central region. Such behaviour is expected if the central region corresponds to the hadronic plateau, discussed in the introduction. In order to quantify the change of $E_T$ with $Q^2$, the value of $E_T$ per unit of pseudorapidity ($E_T$ density) for an $\eta^*$ slice in the central pseudorapidity region ($-0.5<\eta^*<0.5$) and an $\eta^*$ slice in the photon fragmentation region ($-3<\eta^*<-2$) is shown in Fig.~\ref{Q2}. The data clearly demonstrate that the transverse energy in the central region is essentially independent of the $Q^2$ of the photon, but increases significantly in the photon fragmentation region. However, in the photon fragmentation region the difference in the $E_T$ density for DIS and photoproduction data is significant only for large $Q^2$ values, $Q^2 > 20 $ GeV$^2$. The level of transverse energy in the photon fragmentation region for interactions with $Q^2 < 10 $ GeV$^2$ is quite similar, roughly independent of the transverse size, $Q^2$, of the photon. Since the former can be described by perturbative calculations, this observation hints that the result of some predicted features of the final state may be transportable to the hitherto assumed non-perturbative ``VDM region" at $Q^2 = 0$. In Fig.~\ref{Q2} the results are compared with predictions from the CDM and MEPS models, as described in the previous section. The $E_T$ density predicted by the CDM model is found to be independent of $Q^2$, which is in accordance with the data in the central region, but does not reproduce the data in the photon fragmentation region. In this version of the CDM model the scale for QCD radiation is given by the $p_T^2$ of the radiated gluons, which is limited by $W^{4/3}$. Since data corresponding to the same $W^2$ region are selected, the predictions for the $E_T$ density are similar for the different $Q^2$ data samples. A new version of the CDM model (version 4.06)\cite{leif2} partially corrects for this defect in the photon fragmentation region, but still does not yield a good description of the data. The MEPS model on the other hand describes the behaviour of the transverse energy flow in the photon fragmentation region but fails in the central region. Hence, in this model, the influence of the photon virtuality extends too far in rapidity away from the photon-$q\overline{q}$ vertex. The comparison of absolute values of the $E_T$ density between the data and models should be viewed with some caution since diffractive events are not included in these models. These events, which amount to $\approx 6\%$ of all events, give rise to a rapidity gap~\cite{h1rapgap}, i.e. no energy deposition in the central region. The exclusion of diffractive events however does not change the shape of the distribution of the $E_T$ density versus $Q^2$. Comparisons of the energy flows for models and data with rapidity gap events removed can be found in~\cite{h1bfkl}. Within the hadronic picture, the observation of a rise of the $E_T$ density with $Q^2$ in the photon fragmentation region implies that, on arrival at the target, the virtual photon has a hadronic structure (parton configuration) which depends on~$Q^2$. With increasing $Q^2$ the transverse momentum of the constituent partons in this hadronic structure increases. A formal explanation of this effect based on QCD calculations is given in~\cite{khoze}. The central region however shows a more universal behaviour, independent of the partonic nature of the colliding particles. It is therefore interesting to compare our measurement with high energy hadron-hadron collisions. In Fig.~\ref{W} the transverse energy density, $dE_T/d\eta^*$, at $\eta^* = 0$ is shown for DIS, photoproduction, $pp$ and $\overline{p}p$ interactions as a function of $W$ (= $\sqrt{s}$ for hadron-hadron collisions). The results for DIS include only the data with $Q^2 < 50$~GeV$^2$, where enough statistics is available to cover a large $W$ (or equivalently $y$) range, as given in Table 1. The $W$ dependence of $dE_T/d\eta^*$ observed in DIS processes at low $Q^2$ (and low $x$) agrees with the $W$ interpolation between measurements from $p\bar{p}$~\cite{ua1mb} and $pp$~\mbox{\cite{ua1mb,na22}} collisions. Also the photoproduction data show the same level of $E_T$ for the $W$ value of this data sample. This observation is consistent with the ansatz of the analogy of $\gamma^*p$ interactions with real photon-hadron and hadron-hadron collisions. In Fig.~\ref{ET-MAX}a the distribution of (uncorrected) transverse energy summed over a unit of pseudorapidity in the central region ($-0.5<\eta^*<0.5$) is shown for photoproduction and for DIS data ($10<Q^2<100$~GeV$^2$, $0.3<y<0.5$). The comparison shows that in the central region not only the mean $E_T$, but also the energy spectra themselves, are quite similar. This apparently holds down to very low $E_T$ values, where diffractive dissociation processes are expected to dominate (see peak at $E_T=0$). Since for $\eta^* = 0$ essentially the same detector regions are used in DIS and photoproduction interactions the agreement is not influenced by detector effects. Following this observation, it is interesting to compare the fraction of diffractive events in photoproduction and DIS. An effective way to detect diffractive processes at HERA is via the $\eta^*_{max}$ variable, which is the maximum pseudorapidity in the $\mbox{$\gamma^{*}p$}$ CMS frame, of a reconstructed track or calorimetric cluster with an energy larger than 400 MeV observed in the detector. In diffractive events $\eta^*_{max}$ indicates the maximum pseudorapidity of secondary hadrons from photon fragmentation and is related to the fraction $x_p$ of the initial proton momentum carried by the diffractively scattered proton via $\eta^*_{max} \sim \ln(1-x_p) + C$~\cite{kaidalov}. The comparison of the $\eta^*_{max}$ distributions for DIS and photoproduction interactions is shown in Fig.~\ref{ET-MAX}b. The figure shows spectra which fall off rapidly from $\eta^*_{max} = 1$ with decreasing $\eta^*_{max}$. This part of the spectra can be described by non-diffractive DIS and photoproduction interactions as demonstrated in~\cite{h1rapgap,h1diffphoto}. For $\eta^*_{max} \sim -1$ the spectra level off to a constant plateau. The events with this and lower $\eta^*_{max}$ values have been shown to originate predominantly from a diffractive mechanism~\cite{h1rapgap2,h1diffphoto}. Below $\eta^*_{max}<-3 $ the photoproduction and DIS data are affected differently by the detector acceptance and therefore not included in this analysis. As a consequence of this cut the low mass resonance region is avoided as well. For diffractive events the distribution $1/N\cdot dN/d\eta^*_{max} \sim (1/\sigma_{tot})(d\sigma/d(1-x_p))(1-x_p)$ is expected to be roughly independent of $\eta^*_{max}$ if the diffractive cross section $d\sigma/d(1-x_p)$ is approximately inversely proportional to $(1-x_p)$, see ref.~\cite{h1rapgap2} \footnote {It was shown that for DIS events $d\sigma/d(1-x_p)$ is proportional to $(1-x_p)^n$ with $n=1.19\pm0.06\pm0.07$}. This lack of dependence of $1/N\cdot dN/d\eta^*_{max}$ on $\eta^*_{max}$ is itself often taken as the signature of diffraction~\cite{bjdiff}. Fig.~\ref{ET-MAX}b shows clearly that indeed at low $\eta^*_{max}$ both for DIS and photoproduction the differential distribution $1/N\cdot dN/d\eta^*_{max}$ is roughly independent of $\eta^*_{max}$. Furthermore the relative contribution of the photon high mass diffractive dissociation for photoproduction and DIS interactions is found to be the same to within 15-20\%. This agreement is not affected by detector effects since for a given $\eta^*_{max}$ value largely the same detector regions are explored in DIS and photoproduction interactions. This observation is at the same level of agreement as measurements made for hadron-hadron interactions, and is in accord with the expectation of approximate factorization of high mass diffractive cross sections, as observed in hadron collisions and explained with triple Regge phenomenology~\cite{triple}: the ratio of the differential diffractive cross section, $d\sigma/d(1-x_p)$, to the total cross section is approximately independent of the type of dissociating hadron. The fact that this factorization rule also seems to hold for virtual photon interactions gives an additional argument in favour of the validity of a universal hadron-like description of low-$x$ DIS, real photoproduction, and hadron-hadron collisions. \section{Conclusions} The comparison of photoproduction and low-$x$ DIS data at the $ep$ collider HERA reveals striking similarities in the energy flow of the hadronic final state and relative rate of high mass diffractive dissociation of the photon. The $W$ dependence of the transverse energy density in the central rapidity region is found to be similar to that seen in high energy hadron-hadron collisions. These findings are consistent with the hadronic picture of the photon which can therefore be considered to be complementary to the conventional deep-inelastic scattering picture, even at large $Q^2$ values. This picture gives a description of the transition region from the high $Q^2$ perturbative region to the low $Q^2$ non-perturbative region, and provides a basis for a universal description of hadron-hadron, real photon-hadron and virtual photon-hadron high energy interactions. \section*{Acknowledgements} We are grateful to the HERA machine group whose outstanding efforts made this experiment possible. We appreciate the immense effort of the engineers and technicians who constructed and maintained the detector. We thank the funding agencies for their financial support of the experiment. We wish to thank the DESY directorate for the support and hospitality extended to the non-DESY members of the collaboration. We thank further J. Bartels, A. Kaidalov, J. Kwiecinski and M. Ryskin for useful discussions. {\Large\normalsize} \begin{footnotesize}
\section{INTRODUCTION} The visual search for absorption lines in the gamma ray burst spectra accumulated by the Burst and Transient Source Experiment (BATSE) on the {\it Compton Gamma Ray Observatory (CGRO)} has resulted in a number of line candidates, but no definitive detections (Palmer et al. 1994, hereafter Paper~I). Here we present the results of a systematic evaluation of the line candidates identified thus far, and discuss the two most significant candidates. The existence of lines in the BATSE spectra is one of the major unresolved observational issues confronting BATSE since the absorption lines reported by the KONUS (Mazets et al. 1981), {\it HEAO~1} (Hueter 1987) and {\it Ginga} (Murakami et al. 1988) detectors were attributed to cyclotron absorption in $\sim$10$^{12}$ gauss magnetic fields (\cite{wang89}), which suggests that bursts originate on neutron stars, the only known anchors for teragauss fields. This paper is part of a series describing and analyzing our search for lines in the BATSE data. Paper~I reported the absence of line detections by the visual search of the BATSE spectra. Band et al. (1994, hereafter Paper~II) developed the methodology for comparing the absence of a BATSE detection with the detections reported by previous missions; a preliminary calculation shows that the apparent discrepancy between BATSE and {\it Ginga} is not yet compelling. Band et al. (1995, hereafter Paper~III) demonstrated that BATSE would have detected the lines that {\it Ginga} observed; the simulations in Paper~III also provide the probabilities for detecting lines in the BATSE spectra which are necessary for a comprehensive comparison between BATSE and previous detectors. The visual search provides neither the statistics necessary for this comprehensive search, nor measures of its completeness, as is evident from the candidate evaluation presented here. Therefore, we have begun a more sophisticated computerized search (Schaefer et al. 1994). The evaluation of line candidates involves many instrumental and data analysis details which are summarized in \S 2. We first present the results of the candidate evaluation (\S 3), and then discuss two particular candidates in greater detail (\S 3.1-2). \section{METHODOLOGY} We search for line features in the spectra accumulated by BATSE's Spectroscopy Detectors (SDs); the SDs have been described in detail elsewhere (Fishman et al. 1989; Band et al. 1992). In brief, the SDs are simple 5" diameter by 3" thick cylindrical NaI(Tl) scintillation detectors with an energy resolution of 19\% at 60~keV. An SD is found in each of BATSE's eight modules located at the corners of the {\it CGRO} spacecraft. Each module also contains a Large Area Detector (LAD) for detecting transient events, determining their positions, and recording their time histories. For 4-10 minutes after a burst trigger, BATSE accumulates 192 SD spectra with 256 channels spanning two energy decades with a low energy cutoff between 15 and 30~keV (the SHERB data type). These spectra are gathered from four detectors, with a greater number from the most brightly illuminated detectors. The accumulation times are based on the count rate. Background spectra are accumulated every $\sim5$ minutes when BATSE is not in burst mode (the SHER data type). Unfortunately, a few channels just above each spectrum's low energy cutoff are distorted by the SLED ({\bf S}D {\bf L}ow {\bf E}nergy {\bf D}istortion), an electronic artifact discovered after launch (Band et al. 1992). For a typical SD in its highest gain setting the upper end of the SLED is between 15 and 20~keV. Although the SLED can be partially mitigated by the calibration software, the channels affected by the SLED will probably never be trusted for more than determining the continuum below a line candidate. However, for the purpose of establishing consistency among all detectors we do search for features in the SLED-affected channels when lines have been found at the same energy in other detectors. For an observed spectral feature to be considered intrinsic to the incident spectrum (i.e., a ``real'' feature of astrophysical interest), the probability that it is a statistical fluctuation must be small, and the spectra accumulated by all detectors must be consistent. Thus the first detection criterion is that the feature is significant in at least one detector, while the second criterion is consistency among all detectors which could have observed the feature. Note that we do not require multiple independent detections (although it is very rare that a feature would be observable by only one detector), nor must a feature be significant in all detectors where it is observed. As we showed in Paper~III, fluctuations can vary the apparent significance of an absorption line. In addition, the probability that a line will be significant enough to be considered a detection varies with the burst angle (the angle between the burst and the detector normal). Thus we cannot expect significant detections in all detectors. We use the $F$-test to calculate the probability that an observed feature is a fluctuation. As applied here, the $F$-test compares a spectral fit with a continuum-only model to the fit with a continuum+line model. Assume the continuum-only fit gives $\chi_1^2$ with $\nu_1$ degrees-of-freedom, while the continuum+line fit results in $\chi_2^2$ with $\nu_2$. We define the $F$ statistic as (Martin 1971) \begin{equation} F = \left({{\chi_1^2-\chi_2^2}\over{\nu_1-\nu_2}}\right) \left({{\nu_2}\over{\chi_2^2}}\right) \quad . \end{equation} This $F$ statistic is characterized by a normalized distribution which is a function of $\nu_1-\nu_2$ and $\nu_2$; the integral over this distribution from $F$ to infinity is the $F$-test probability $P(\ge F)$, which we define as the significance of the feature, with a small probability indicating a more significant feature. We choose a maximum probability $P(\ge F)$ as the threshold for a candidate to be considered a detection. Note that $P(\ge F)$ is only the {\it a priori} probability of a fluctuation with the observed parameters; it does not consider all the possible ``trials'', the range of parameters (e.g., time intervals, line centroids, widths, etc.) in which line-like fluctuations could have occurred. If there are a large number of trials then it is likely that in some trial there will be a rare fluctuation. The probability that a fluctuation will produce a spurious line feature with a given significance somewhere in our dataset is the product of $P(\ge F)$ and the number of trials; therefore we require a very small $P(\ge F)$ to conclude the line is real. The calculation of the number of trials in our data is a difficult issue in standard ``frequentist'' statistics which we will address later in this series of papers. We estimate there are of order $\sim 3000$ trials in our spectra, and therefore use a detection threshold of $P(\ge F)\le 10^{-4}$. While this threshold may not appear stringent enough since the probability of a false positive is then $\sim30\%$, we also require consistency among detectors (Paper~I). The expected distribution of fluctuations will be analyzed in greater detail later in this series of papers. Only $\sim 20\%$ of the bursts detected by BATSE have a sufficiently high count rate to warrant detailed spectral analysis, which roughly corresponds to a LAD peak count rate over 10,000 counts-s$^{-1}$ in the 50-300~keV band (i.e., LAD discriminators 2 and 3). Thus of the over 1000 bursts detected, only $\sim$250 have been searched, and of these bursts only $\sim$50 were intense enough for lines comparable to those observed by \hbox{\it Ginga}\ to be detectable. Although the spacecraft telemetry provides SHERB spectra for the four most brightly illuminated detectors, not all the data are appropriate for line searches. First, the detectors must be run at high gain to extend the accumulated spectra low enough to cover the energy range where previous missions reported spectral features. For most of the mission, two of the eight SDs have been operated at low gain (to satisfy other mission objectives), and the gain of one of the nominally high gain detectors cannot be pushed high enough to cover energies below $\sim30$~keV. Second, the burst angle may be greater than $\sim 80^\circ$, the angle at which the spacecraft and the rest of the BATSE module affects the incident flux and our detector response model becomes unreliable. The SHERB data are from the SDs in the modules of the four most brightly-illuminated LADs; these LADs normally have burst angles less than $90^\circ$. However, the detector normals of the SDs and LADs are offset by $18.5^\circ$, and therefore a module whose LAD has a burst angle less than $90^\circ$ may have an SD with an angle greater than $90^\circ$. Earth-scatter of GRB photons off the atmosphere only complicates matters by occasionally increasing the count rate in detectors with large burst angles. Thus usable spectra may be available from only two or three SDs. The visual search inspects background-subtracted spectra on different time scales. By adding together the basic SHERB spectra, we construct spectra for the entire burst and for the major temporal substructures (e.g., each major emission spike). In addition, spectra are produced on the shortest time scale over which all the detectors are accumulated, that of the third and fourth rank detectors (i.e., the SDs in the third and fourth most brightly illuminated modules). Thus spectra are available from all relevant detectors over the same time ranges, allowing comparisons on the same time scale. The resulting background-subtracted spectra are inspected visually for both emission and absorption lines over the entire energy range. Features which appear to be very significant are studied further by inspecting spectra averaged over different time ranges. Thus a feature might be identified in a spectrum averaged over the third rank accumulation time (e.g., from 3.328 to 6.592~s after the trigger) but might be most significant in a spectrum averaged over an overlapping time range (e.g., from 4.160 to 7.360~s). This search is sensitive to features which are significant on the searched time scales, or which are apparent to the eye in the searched time intervals but become more significant by varying the time range; thus this search is not sensitive to short lived features which are washed out in spectra accumulated over longer periods, nor to features which persist a long time but are not evident in the searched spectra over shorter time periods. Our simulations indicate that the eye is sensitive to features with an $F$-test significance of $10^{-2}$ or smaller (i.e., $P(\ge F)\le 10^{-2}$). Since in this study we find that we identified many candidates with significances of order 0.1, we should have detected features which persisted over many of the searched spectra. We evaluate the line candidates identified by the visual search. Because the gain differs between detectors and varies with time, each burst's spectra must be calibrated for each detector (Band et al. 1992). Our response matrices include Earth-scatter. Background spectra are created channel-by-channel by fitting 2nd, 3rd, or 4th order polynomials to spectra before and after the burst. As always, background creation is something of an art; we evaluate the calculated background by the quality of the fits to each channel. These fits provide the background as a function of time; for background subtraction these functions are integrated over the time interval of interest. We fit the spectra with the standard Marquardt-Levenberg algorithm minimizing $\chi^2$ (Bevington 1969, pp.~232-241; Press et al. 1992, pp.~678-683). In brief, the parameters of a model photon spectrum are varied to minimize the difference---quantified by $\chi^2$---between the calculated count spectrum (the model photon spectrum folded through the detector response) and the observed spectrum. We define $\chi^2$ with model variances, requiring additional terms in the gradients used by the Marquardt-Levenberg algorithm (see the appendix of Ford et al. 1995). The continuum is usually fit with the four parameter ``GRB'' (for lack of a better name) model (Band et al. 1993), \begin{eqnarray} N_C(E) =& A (E/100)^\alpha \exp\left[-\left({E\over{E_0}}\right)\right]\quad , \quad &E\le (\alpha-\beta)E_0 \\ =& A^\prime (E/100)^\beta \qquad , \quad &E> (\alpha-\beta)E_0 \quad ,\nonumber \end{eqnarray} where $A^\prime$ is chosen so that $N_C(E)$ is continuous at $E=(\alpha-\beta)E_0$. In a few cases where we find $E_0$ to be very large or $\beta$ is nearly equal to $\alpha$, we simplify the continuum model, either by using a simple power law or by dropping the high energy power law in eq.~2 (the COMP model); the decision to simplify the continuum can be justified in terms of the effect on the reduced $\chi^2$. An apparent absorption line is usually modeled as a multiplicative line, \begin{equation} F(E)=N_C(E)\, \exp\left(-\phi\left(E\right)\right) \quad , \end{equation} and an emission line as an additive line, \begin{eqnarray} F(E)=N_C(E) + \phi(E) \quad , \end{eqnarray} where $\phi(E)$ is the Gaussian line profile (with three parameters: amplitude, line centroid and line width). For absorption lines whose intrinsic width appears to be much narrower than the instrumental resolution we also use a simple two parameter ``black'' line model with no flux over an energy range centered on the line centroid; for very narrow lines the instrument is unable to resolve the line profile. We always fit the spectra over the greatest energy range possible. The low energy cutoff is chosen to avoid the SLED while the highest available energy channel is included. In Paper~III we found that extending the fitted energy range to energies much greater than the features of interest increased the significance of the features, in part because the continuum is better determined. On the other hand, we were concerned that the choice of the low energy cutoff $E_1$ could make a large difference in the calculated significances. The exact value of $E_1$ which should be used is not automatic because the upper end of the SLED artifact is not clear-cut, and analysts may choose different values (in particular there is a tendency to choose round numbers, e.g., 20~keV when the SLED ends at 17~keV). To test the effect of varying $E_1$, we calculated the significance of the 54~keV candidate absorption line in GB~920307 using different values of $E_1$, as is shown by Figure~1. As can be seen, there is not a meaningful change in significance for $E_1$ between 15 and 30~keV, and the significance decreases ($P(\ge F)$ increases) only as $E_1$ approaches the candidate line energy. Thus small changes in $E_1$ should not make a large difference. Of course, the significance will be sensitive to the choice of $E_1$ if the line energy is near the low energy cutoff. \section{RESULTS} Table~1 describes the candidates and the fits to them, while Figure~2 presents the distribution of their significances. Few candidates are more significant than $P(\ge F)\le 0.01$, and none exceeds our detection threshold of $P(\ge F) = 10^{-4}$, although two nearly satisfy the significance criterion; these two are discussed in greater detail below. The significance distribution of the line candidates in Figure~2 is shaped by a number of components. First are the inevitable line-like statistical fluctuations, with many low significance fluctuations and few at the high end (i.e., small values of $P(\ge F)$). The normalization of this distribution is proportional to the number of trials in the burst dataset, a poorly determined quantity. Second is the distribution produced by real spectral lines (if lines exist). These lines are characterized by a (currently unknown) distribution of parameters. Paper~III showed that as a result of the fluctuations in any given observation, the same line may be observed by identical detectors with very different significances. Thus the observed significance distribution is the convolution of the line parameter distribution and the distribution of significances for a given set of line parameters. Note that it is entirely possible that many of our low significance (large $P(\ge F)$ value) candidates which do not satisfy our detection criteria may be real lines. Finally, incompleteness truncates the low significance end of the observed significance distribution; most candidates with $P(\ge F)\sim 0.1$ will not be identified by the search. We suspect that most or all of our line candidates are fluctuations. Our intention was to identify those features which appeared to have $P(\ge F)\le 0.01$; the large number of candidates with $P(\ge F)\ge 0.1$ leads us to suspect the accuracy of the qualitative (i.e., ``eyeball'') assessment of the significance of a line. We would expect equal numbers of emission line-like and absorption line-like fluctuations. However, absorption features were more readily recorded because of the previous reports of absorption lines in the 15-100~keV range (as discussed in Paper~II), although emission features up to 511~keV were also flagged (emission lines have also been reported by previous missions). In addition, absorption lines may be more apparent to the eye than emission lines of comparable strength because of the curvature of the burst count spectra. Thus we find that the number of absorption line candidates (67) exceeds that of emission line candidates (13). The visual search does not provide the statistics of the lines which could have been detected in the searched spectra. These statistics are necessary to compare different missions (e.g., BATSE with \hbox{\it Ginga}) and to place constraints on the frequency with which different types of lines occur (Paper~II). In addition, the search does not produce measures of its completeness (e.g., the fraction of features with $P(\ge F)\sim 10^{-2}$ which were identified by the search). Since the visual search relies on human judgement which can be affected by fatigue and distractions, there is always the fear that an interesting candidate may have been missed. While we are confident that we would have detected features with $P(\ge F)\le 10^{-4}$ in the inspected spectra, we are relying on identifying less significant features in order to detect lines which persist over many of the searched spectra. Thus we cannot conclude definitively that detectable lines do not exist in the BATSE spectra. For these reasons we have begun a computerized search (Schaefer et al. 1994) which is unaffected by human subjectivity and which will provide relevant statistics. In addition, this computerized search will consider almost all possible spectra which can be formed from consecutive SHERB spectra. \subsection{The Feature in GB~920315} The most significant candidate absorption line is found in GB~920315 with $P(\ge F) = 1.6 \times 10^{-4}$, just above our detection threshold. The candidate is a feature at 80~keV in a spectrum accumulated between 1.152 and 1.728~s from SD1 (i.e., the SD in module 1), which is the 2nd rank detector with a burst angle of 58.3$^\circ$. The 1st rank detector, SD3, was in a low gain setting (and therefore did not cover 80~keV), and the 3rd and 4th rank detectors, SD6 and SD7, had burst angles of 121.7$^\circ$ and 105.4$^\circ$, respectively. Therefore, SD1 was the only SD capable of seeing a feature below $\sim 300$~keV, making consistency between detectors automatic; had the feature in SD1 met our significance criterion, it would have technically been a detection. Features are not evident at 80~keV in LAD spectra (with poorer spectral resolution but better statistics than SD spectra), although the sensitivity of the LADs to lines of this type has not yet been determined. As can be seen from the count spectrum in Figure~3, the candidate at 80~keV draws much of its significance from one particularly low channel; however, this low channel is surrounded by a number of channels which dip below the general trend. As we argued above, there is a probability of order $\sim 30$\% of a fluctuation with a significance of $\sim 10^{-4}$ somewhere within our data set, and therefore this feature can be attributed to a statistical fluctuation. Indeed, we would not have been comfortable declaring this candidate a detection had its significance been a factor of 1.6 smaller because of the absence of confirmation by other detectors and the importance of one low point. Our first definitive detection must be more convincing. \subsection{The Feature in GB~930506} The candidate in GB~930506 has attracted a great deal of attention (Ford et al. 1994; Preece et al. 1994; Freeman \& Lamb 1994; Paper~I). An absorption feature was found at 55 keV in SD2; Figure~4 presents the count spectrum. The feature has been evaluated with fits by various continuum and line models over different energy ranges; in many cases the resulting $P(\ge F)$ satisfied our detection criterion. However, the feature in SD2 is absent in the spectra from other BATSE detectors. Two other SDs and two LADs, all with smaller burst angles than SD2, also viewed the burst (see Table~2). Unfortunately, the SD with the smallest burst angle (SD3) was at a low gain setting, and therefore did not cover the energy of the line candidate. In SD7 the upper edge of the SLED fell at the energy of the line in SD2. Using the methodology of Band et al. (1992), Ford et al. (1994) developed an {\it ad hoc} correction for the SLED in SD7. With this correction, no line is evident at 55~keV in SD7; fitting the SD7 spectrum with a continuum+line model found a weak line at 55.9~keV with $P(\ge F)=0.73$. The probability that the line observed in SD2 (i.e, with the parameters fit to the SD2 spectrum) would be as nearly unobservable as the line fit to SD7 is less than one percent; specifically, none of 200 simulated SD7 spectra using the parameters of the SD2 fit resulted in $P(\ge F)$ as large as 0.73. Similarly, a joint fit to SD2 and SD7 does not find a significant line feature (Briggs 1995). No lines are evident in the two LADs with small burst angles (Preece et al. 1994). We reanalyzed the feature in SD2 with the procedures used to evaluate the other line candidates; the results are presented by Table~3. We tried three different continuum models---a power law, COMP and GRB---and two different line models---multiplicative and ``black.'' The background varied significantly over the $\sim 2000$~s before and after the burst, and there were a number of reasonable choices of background spectra from which we could interpolate the background at the time of the candidate. Consequently we calculated the line significances for the six continuum+line models (three continuum and two line models) for two different interpolated backgrounds; as can be seen from Table~3, the resulting line significances are qualitatively similar using the two background spectra. From the reduced $\chi^2_\nu$ for the different fits, also presented in Table~3, it is clear that the power law fits are poor, while the COMP and GRB fits are acceptable ($\chi^2_\nu\sim 1$). While the power law fits are highly significant, $P(\ge F)\sim 10^{-7}$, the COMP and GRB fits do not meet our significance criterion for a detection. It is our experience that when a continuum model does not fit a spectrum well, the line model assists in fitting the observed continuum; hence a poor continuum model results in a significant line fit. Thus we conclude that the high significance of the power law fits is not an accurate representation of the line significance. The fitted line is narrower than the instrumental resolution (a resolution of $\sim 11$~keV as opposed to a line width of $\sim3.9$~keV), and therefore using the ``black'' line model, with one fewer parameter than the multiplicative line model, is more significant by a factor of $\sim 4$. Nonetheless, the line does not satisfy our significance criterion for any of the acceptable line fits. In their extensive analysis of this candidate, Freeman \& Lamb (1994) performed fits to SD2 and joint fits to SD2+SD7, SD2+SD3, and SD2+SD3+SD7 using both the COMP and GRB continua. The significances of the joint fits to SD2+SD3 are $1.6\times 10^{-5}$ and $7.4\times 10^{-5}$ for the COMP and GRB continua, respectively. We do not believe this affects our conclusion that the feature is not significant. First, the relevant significance is the largest $F$-test probability that results from a reasonable continuum model; to show the feature is significant we must prove that it cannot be produced by a reasonable continuum shape. Thus we should use the significance from the joint fit with the GRB continuum which just barely satisfies our first criterion. Second, adding a second detector doubles the number of degrees-of-freedom. As discussed in Paper~III, extending the energy range of a fit can improve the $F$-test probability both by better determining the continuum, thereby increasing $\Delta \chi^2$ between the continuum and continuum+line fits, and by increasing the degrees-of-freedom. The first effect is a true reflection of the line significance, while the second effect is an artifact of the $F$-test. We found in Paper~III that a factor of two increase in the number of degrees-of-freedom decreased the $F$-test probability by an order of magnitude; therefore we conclude that most of the improvement of the significance of the SD2+SD3 joint fit over the SD2 fit results from this degrees-of-freedom effect. We therefore conclude that the line candidate in GB~930506 meets neither of our detection criteria. The $F$-test probability for the line feature in SD2 is small but not so small as to lead us to reject the possibility that the feature is a rare statistical fluctuation. \section{SUMMARY} We have evaluated the line candidates identified by the still-ongoing visual search of burst spectra observed by the BATSE SDs. We find that none of the candidates satisfy our detection criteria; indeed most of the candidates are not very significant. Taken at face value, this evaluation of line candidates from the visual search leaves us with no BATSE line detections, while KONUS, {\it HEAO~1} and \hbox{\it Ginga}\ all reported detections. However, our approximate assessment of the consistency between BATSE and \hbox{\it Ginga}\ (Paper~II), accurate to a factor of $\sim$2, indicates that the apparent discrepancy between these two missions is not yet compelling; a more definitive calculation is in progress. Thus based on the BATSE observations there is no reason yet to question the detections reported by previous missions. Clearly, we will be forced to conclude that BATSE is discrepant with previous missions if no lines are detected after many additional intense BATSE bursts are observed. The visual search has not provided us with measures of its sensitivity; we do not know what part of the space of line parameters has been probed. In addition, the visual search is affected by the subjectivity of the human eye, and there is always the fear that the search may have missed a significant line. Consequently we have begun a new computerized search which should overcome these limitations. Also, this search will consider progressively longer accumulations, thereby increasing the range of lines which can be found. Both the ongoing visual search and this computerized search will undoubtedly generate new candidates, and perhaps ultimately a few detections. \acknowledgments We thank P.~Freeman and D.~Lamb for insightful discussions regarding GB~930506. The work of the UCSD group is supported by NASA contract NAS8-36081. \clearpage \addtocounter{page}{2} \setcounter{table}{1} \begin{table*} \begin{center} \smallskip \begin{tabular}{l c c} \tableline \tableline Detector & Burst Angle\tablenotemark{a} & Energy Range (keV)\tablenotemark{b} \\ \tableline SD3 & 43.7$^\circ$ & 250-25000 \\ SD7 & 56.2$^\circ$ & 55-4800 \\ SD2 & 73.7$^\circ$ & 15-1350 \\ LAD3 & 28.8$^\circ$ & 35-1800 \\ LAD7 & 42.8$^\circ$ & 35-1800 \\ \end{tabular} \end{center} \caption{Detectors viewing the GB~930506 candidate} \tablenotetext{a}{The angle between the detector normal and the burst.} \tablenotetext{b}{The energy range covered by the detector.} \end{table*} \clearpage \begin{table*} \begin{center} \begin{tabular}{l c c c c c} \tableline \tableline \multicolumn{2}{c}{Continuum Model} & \multicolumn{2}{c}{Black Line\tablenotemark{a}} & \multicolumn{2}{c}{Multiplicative Line\tablenotemark{b}} \\ Type\tablenotemark{c} & $\chi^2_\nu$ & $P(\ge F)$ & $\chi^2_\nu$ & $P(\ge F)$ & $\chi^2_\nu$ \\ \tableline PL & 1.96688 & $1.28\times10^{-7}$ & 1.72630 & $6.62\times10^{-7}$ & 1.73506 \\ & 1.97315 & $1.04\times10^{-7}$ & 1.72861 & $2.85\times10^{-7}$ & 1.72738 \\ \tableline COMP & 1.00083 & $2.46\times10^{-4}$ & 0.93837 & $9.30\times10^{-4}$ & 0.94320 \\ & 0.99953 & $1.87\times10^{-4}$ & 0.93489 & $7.98\times10^{-4}$ & 0.94063 \\ \tableline GRB & 0.99894 & $3.20\times10^{-4}$ & 0.93850 & $1.18\times10^{-3}$ & 0.94324 \\ & 0.99772 & $2.41\times10^{-4}$ & 0.93502 & $9.74\times10^{-4}$ & 0.94043 \\ \end{tabular} \end{center} \caption{Significance of GB~930506 candidate in SD2} \tablenotetext{a}{A line model with a rectangular profile centered on the line centroid (2 parameters).} \tablenotetext{b}{The line model is an exponential of a Gaussian line profile (eqn.~3---3 line parameters).} \tablenotetext{c}{The type of continuum model: PL---power law (2 parameters); COMP---power law with exponential break (3 parameters); GRB---power law with exponential break flattening to high energy power law (eqn.~2---4 parameters). Note that PL is a subset of COMP, which in turn is a subset of GRB.} \tablecomments{Fits to the line candidate at 55~keV in the SD2 spectrum accumulated over 6.080-9.920~s. There are two entries for each case corresponding to two different background spectra. The spectrum was fit over 232 channels covering 15.3--1351~keV.} \end{table*} \clearpage
\section{Introduction} \ Instanton-induced processes in the standard electroweak theory are known to lead to baryon and lepton number violation. Although 't Hooft \cite{h1} showed several years ago that such phenomena are utterly suppressed by the factor $e^{-8\pi ^2/g^2}$ ($g$ is the $SU(2)_L$ gauge coupling constant), several authors \cite{c2} explored the possibility that this exponential suppression factor can be overcome at high energies by the phase space which corresponds to multiparticle production. The key observation is that the $% SU(2)_L$- instanton induces to leading semiclassical approximation effective point-like interactions which involve all the fermionic left-handed $SU(2)_L$% -doublets of the theory (four per generation) and any number of Higgs and gauge bosons. The inclusive cross section of the baryon and lepton number violating two fermion scattering can then be calculated as the imaginary part of the forward $2\rightarrow 2$ scattering amplitude depicted in fig.1 \cite{p3}. As a result, this inclusive cross section appears to grow exponentially with energy and can conceivably become unsuppressed at energies of the order of the sphaleron \cite{m4} mass. This leading order behavior may, however, be drastically altered by higher order corrections well before the energy reaches the sphaleron mass and consequently, these phenomena may remain unobservable at all energies. Several authors \cite{a5} actually, have suggested that this could be the case if multiinstanton corrections were to be taken into account. Corrections to the $2\rightarrow 2$ scattering amplitude consisting of linear instanton-antiinstanton chains in alternating order were considered, with particle exchange allowed only between successive instantons and antiinstantons. Dorey and Mattis \cite{d6}, however, using the valley method \cite{k7}, pointed out that inclusion of non-nearest neighbor instanton-antiinstanton as well as instanton-instanton and antiinstanton-antiinstanton interactions could render these chain graphs unimportant at the relevant energies. This, however, may not happen if Dorey's result on the non linear $O(3)$ $\sigma -$model applies in the realistic case too. The imaginary part of the chain graphs being ultraviolately divergent, requires the introduction of an appropriate cut-off. Finally, these graphs do not include initial state corrections and thus, are not expected to alter the high energy behavior of the leading semiclassical approximation \cite{k7}% ,\cite{m8}. In this work we choose to deal with a class of ladder graphs shown in fig.2. The imaginary part of these graphs turns out to be finite and can, in principle, be unambiguously calculated. They can, in some sense, be thought as including initial state corrections too since the incoming particles enter in different instanton vertices. In addition, such ladder graphs are known to dominate the high-energy behavior in ordinary field theories.We should emphasize, however, that including the ladder graphs of fig.2 does not solve the problem of initial state corrections. Indeed, the entire picture of separated instanton-antiinstanton chains is an uncontrolled approximation at energies where such chains actually become important. \section{The leading semiclassical approximation.} Consider the inclusive cross section, $\sigma _{inc}$, of the B and L violating reaction \begin{equation} \label{eq1}q+q\rightarrow (3n_g-2)\overline{q}+n_g\overline{l}% +\,any\,\#\,of\,Higgses, \end{equation} where $n_g$ is the number of fermion generations ($n_g\geq 1$), $q$ and $l$ represent quarks and leptons respectively and we have ignored for simplicity the production of gauge particles. In the leading instanton approximation, $% \sigma _{inc}$ can be determined by first calculating the forward $% 2\rightarrow 2$ scattering amplitude in Euclidean space as shown in fig.1. Then, $\sigma _{inc}$ is given by the imaginary part acquired by this amplitude after rotating the total incoming Euclidean particle momentum $% p=p_1+p_2$ to Minkowski space $\ (p^2\rightarrow e^{-\imath \pi }\,p^2)$. The expression corresponding to the Euclidean space forward scattering amplitude is \cite{c2},\cite{p3} \begin{equation} \label{eq2}C\int_0^\infty d\rho ^2\,\rho ^{2\alpha }\,e^{-\pi ^2v^2\rho ^2}\,\int_0^\infty d\tilde \rho ^2\,\tilde \rho ^{2\alpha }\,e^{-\pi ^2v^2% \tilde \rho ^2}\,\int d^4x\,e^{-\imath px}\,F(x^2,\rho ^2,\tilde \rho ^2). \end{equation} $C$ is a constant given by \begin{equation} \label{eq3}C\propto \frac 1{\pi ^2}\,(32\pi ^2)^{4n_g-1}\,\left( \frac{8\pi ^2}{g^2}\right) ^8\,\mu ^{\frac{43-8n_g}3}\,e^{-\frac{16\pi ^2}{g^2(\mu )}% },\ \alpha =(28n_g+7)/12, \end{equation} $\rho $ and $\tilde \rho $ are the scale sizes of the instanton and the antiinstanton, \thinspace $x_\mu $ is their Euclidean separation, $v=246GeV$ the electroweak breaking scale, g the gauge coupling and $\mu $ is the renormalization point. The function $F(x^2,\rho ^2,\tilde \rho ^2)$, to leading semiclassical order (or for $x^2\rightarrow \infty $), can be written as \begin{equation} \label{eq4}F(x^2,\rho ^2,\tilde \rho ^2)\equiv F(x^2)=e^{\kappa /x^2}\,\frac 1{(x^2)^n}, \end{equation} where $\kappa =\pi ^2\rho ^2\tilde \rho ^2v^2$ and $n=3(2n_g-1)\geq 3$. The exponential factor in the formula above, corresponds to the Higgses in the final state of reaction (1), whereas the second factor corresponds to the $% 4n_g-2$ fermions which are also being produced. Since we are interested in the high energy behavior of $\sigma _{inc}$, we can assume throughout this work that all fermions and Higgs bosons are effectively massless. We will first consider the integral over the instanton-antiinstanton separation \begin{equation} \label{eq5}I^0(p^2)\equiv \int d^4x\,e^{-\imath px}\,F(x^2)\quad . \end{equation} This integral converges at infinity for $n\geq 3$ , but diverges badly at $% x=0$. This virulent ultraviolet divergence in Euclidean space is due to the attractive nature of the Coulomb potential $V(x^2)=-\kappa /x^2$ resulting from Higgs particle exchange between the instanton and the antiinstanton and is an artifact of the leading semiclassical approximation. We, thus, define a regularized integral \begin{equation} \label{eq6}I_\delta ^0(p^2)=\int_{x^2\geq \delta ^2}\,d^4x\,e^{-\imath px}\,F(x^2) \end{equation} by removing from the range of integration a four-dimensional disc of finite radius $\delta >0$ centered at the origin. Performing the angular integrations, we obtain \begin{equation} \label{eq7}I_\delta ^0(p^2)=2\pi ^2\int_\delta ^\infty dr\,e^{\kappa /r^2}\,r^{3-2n}\,G_{0\,2}^{1\,0}(\frac{p^2r^2}4|0,-1), \end{equation} \noindent where \begin{equation} \label{eq8}G_{0\,2}^{1\,0}(\frac{p^2r^2}4|0,-1)=\frac 1{2\pi \imath }% \,\int_Ldz\left( \frac{p^2r^2}4\right) ^z\frac{\Gamma (-z)}{\Gamma (2+z)} \end{equation} is the well-known Meijer function. The contour $L$ is a loop starting and ending at $+\infty $ and encircling the poles of $\Gamma (-z)$ once in the negative direction. Since $p^2r^2/4$ is positive, one can show that the contour $L$ can be distorted to become parallel to the imaginary axis and lying in the strip $-1/2<Re(z)<0$. Then substituting eq.(8) in eq.(7) and interchanging the order of integrations, we get \begin{equation} \label{eq9}I_\delta ^0(p^2)=\frac 1{2\pi \imath }\,\int_{c-\imath \infty }^{c+\imath \infty }\,dz\,\left( \frac{p^2}4\right) ^z\,\frac{\Gamma (-z)}{% \Gamma (2+z)}\,\pi ^2\,(-\kappa )^{z-n+2}\,\gamma (n-z-2,-\kappa /\delta ^2), \end{equation} $$ -1<c<0. $$ Here, $\gamma (\alpha ,x)$ is the incomplete gamma function which can be expressed as \begin{equation} \label{eq10}\gamma (\alpha ,x)=x^\alpha \,\Gamma (\alpha )\,\gamma ^{*}(\alpha ,x)\ , \end{equation} with $\gamma ^{*}(\alpha ,x)$ being an analytic function of $\alpha $ and $x$% .Eq.(9) then becomes \begin{equation} \label{eq11}I_\delta ^0(p^2)=\frac{\pi ^2\delta ^{4-2n}}{2\pi \imath }% \,\int_{c-\imath \infty }^{c+\imath \infty }\,dz\left( \frac{p^2\delta ^2}4% \right) ^z\,\frac{\Gamma (-z)\,\Gamma (n-2-z)}{\Gamma (2+z)}\,\gamma ^{*}(n-z-2,-\kappa /{\delta ^2}), \end{equation} $$ -1<c<0. $$ Notice, that after interchanging the order of integrations, the range of $c$ can be extended.The imaginary part acquired by $I_\delta ^0(p^2)$ after rotating $p^2$ to Minkowski space $(p^2\rightarrow e^{-\imath \pi }p^2)$, comes from the $\ln p^2$ terms in the series expansion of the right-hand side of eq.(11). These terms are produced by the double poles of the integrand in eq.(11) at $z=m$,$~m=n-2,n-1,n,\ldots $ \nolinebreak. The result is $$ ImI_\delta ^0(e^{-\imath \pi }p^2)=\pi ^3\delta ^{4-2n}\sum_{m=n-2}^\infty \left( \frac{p^2\delta ^2}4\right) ^m\,\frac{(-1)^{m-n+2}}{% m!\,(m+1)!\,(m-n+2)!}\,\times $$ \begin{equation} \label{eq12}\gamma ^{*}(n-m-2,-\kappa /{\delta }^2). \end{equation} Using eq.(10), one can show that \begin{equation} \label{eq13}\gamma ^{*}(-\alpha ,x)=x^\alpha ,\quad \alpha =0,1,2,\ldots , \end{equation} which implies \begin{equation} \label{eq14}ImI_\delta ^0(e^{-\imath \pi }p^2)=\pi ^3{\kappa }% ^{2-n}\,\sum_{m=n-2}^\infty \left( \frac{p^2\kappa }4\right) ^m\,\frac 1{% m!\,(m+1)!\,(m-n+2)!}\,\cdot \end{equation} We now perform the integrals over the sizes of the instanton and the antiinstanton to get the well-known semiclassical result. \begin{equation} \label{eq15}\sigma _{inc}^0=\frac{\pi ^3C}{(\pi ^2v^2)^{4+2\alpha -n}}% \,\sum_{m=n-2}^\infty \left( \frac{p^2}{4\pi ^2v^2}\right) ^m\,\frac{[\Gamma (\alpha +m-n+3)]^2}{m!\,(m+1)!\,(m-n+2)!}\,\ \cdot \end{equation} It is important to note that the $\delta $-dependence of the imaginary part of $I_\delta ^0(e^{-\imath \pi }p^2)$ has completely disappeared in eq.(15). Consequently, $\sigma _{inc}^0$ is $\delta $-independent for any $\delta >0$ and its exponential growth with energy results only from the boundary at infinity of the Euclidean $x$-space in eq.(6).The virulent ultraviolate divergence, as well as the contribution of any ''finite'' part of the Euclidean $x$-space in eq.(6), do not seem to play any essential role. \section{The ladder graphs.} We will now turn to the calculation of the Euclidean space ladder graphs shown in fig.2. The graphs, after rotating $p^2\rightarrow e^{-\imath \pi }\,p^2$ and taking the imaginary part, constitute an important class of multiinstanton corrections to the leading semiclassical approximation of $% \sigma _{inc}$. The forward scattering amplitude which corresponds to the ladder graph with $k+1$ rungs $(k=0,2,4,\ldots )$ is given by \begin{equation} \label{eq16}C^{k+1}\,\prod_{\imath =0}^k\,\prod_{\jmath =1}^k\,\int_0^\infty \,d\rho _\imath ^2\,\rho _\imath ^{2\alpha }\,e^{-\pi ^2v^2\rho _\imath ^2}\,\int_0^\infty \,d\widetilde{\rho _\imath }^2\,\widetilde{\rho _\imath }% ^{2\alpha }\,e^{-\pi ^2v^2\widetilde{\rho }_\imath ^2}\times \end{equation} $$ \int \,d^4q_\jmath \,\frac 1{q_\jmath ^2}\,\int \,d^4x_\imath \,e^{-\imath (q_{\imath }+q_{\imath +1})x_\imath }\,F(x_\imath ^2,\rho _\imath ^2, \widetilde{\rho }_\imath ^2)\quad , $$ where $q_0=p_1$, $q_{k+1}=p_2$ and the four momenta $q_\jmath \,(\jmath =1,2,\ldots ,k)$ are indicated in fig.2, $\rho _\imath \,,\widetilde{\rho _\imath }\,\,\,(\imath =0,1,\ldots ,k)$ are the scale sizes of the $i$-th instanton and the $i$-th antiinstanton respectively and $x_\imath $ is their Euclidean separation. We will first consider the integrals over the instanton-antiinstanton separations \begin{equation} \label{eq17}I_\delta ^k(p^2)=\prod_{\jmath =1}^k\,\int \,d^4q_\jmath \,\frac 1{q_\jmath ^2}\,\,\prod_{\imath =0}^k\,\int\limits_{x_\imath ^2\geq \delta ^2}\,d^4x_\imath \,e^{-\imath (q_\imath +q_{\imath +1})}\,F(x_\imath ^2,\rho _\imath ^2,\widetilde{\rho }_\imath ^2)\quad , \end{equation} where the Euclidean space ultraviolet divergences are again regularized by restricting the $x_i$-integrations to $x_\imath ^2\geq \delta ^2$. Repeating the analysis of the previous section, we obtain\ $$ I_\delta ^k(p^2)=\frac{\pi ^{2(k+1)}\,\delta ^{(4-2n)(k+1)}}{(2\pi \,i)^{k+1}% }\,\prod_{\imath =0}^k\,\int_{c_\imath -\imath \infty }^{c_\imath +\imath \infty }\,dz_\imath \,\left( \frac{\delta ^2}4\right) ^{z_\imath }\,\frac{% \Gamma (-z_\imath )\,\Gamma (n-2-z_\imath )}{\Gamma (2+z_\imath )}\times $$ \begin{equation} \label{eq18}\gamma ^{*}(n-z_\imath -2,-\frac{\kappa _\imath }{\delta ^2}% )\,\,A_k(z_0,z_1,\ldots ,z_k) \end{equation} where $\kappa _\imath =\pi ^2\rho _\imath ^2\,\widetilde{\rho }_\imath ^2\,v^2$ and (see fig.2) \begin{equation} \label{eq19}A_k(z_0,z_1,\ldots ,z_k)=\prod_{\jmath =1}^k\,\int \,d^4q_\jmath \,\frac 1{q_\jmath ^2}\,\prod_{\imath =0}^k\,(q_\imath +q_{\imath +1})^{2z_\imath }\quad . \end{equation} The constants $c_\imath \ (\imath =0,1,\ldots ,k)$ which satisfy the inequalities $-1<c_\imath <\nolinebreak0\,$(see eq.(11)) may and in fact will have to be further restricted in eq.(18) so that $A_k(z_0,z_1,\ldots ,z_k)$ and the $z_\imath $ -integrals exist. For the moment we just assume that there is some region of $z_\imath $ 's in which $A_k(z_0,z_1,\ldots ,z_k)$ exist and we restrict ourselves in this region. This assumption will be proved to be correct a posteriori (see below). The $s=(p_1+p_2)^2$ dependence of $A_k$ at $p_1^2=p_2^2=0$ can be easily found from dimensional arguments (there are no infrared divergences)\footnote{% This semieuristic argument can be further corroborated by an explicit although tedious calculation.}. We get \begin{equation} \label{eq20}A_k(z_0,z_1,\ldots ,z_k)=F_k(z_i)\,s^{k+\sum_i\,z_i}. \end{equation} It is easily seen that, for $p_1^2=p_2^2=0,$we also have \begin{equation} \label{eq21}p_2\cdot \frac \partial {\partial p_2}A_k=s\,\frac \partial {% \partial s}A_k, \end{equation} and, thus, \begin{equation} \label{eq22}p_2\cdot \frac \partial {\partial p_2}A_k=(k+\sum_i\,z_i)A_k=z_k% \,A_k-z_k\,A_k^{\,q_k}(z_0,\,z_1,\ldots ,z_k-1). \end{equation} \noindent $A_k^{\,q_k}$ denotes the expression $A_k$ with the $q_k$- propagator ommited. Eq.(22 ) then gives \begin{equation} \label{eq23}A_k=-\frac{z_k}{k+\sum_{i\neq k}\,z_i}\,A_k^{\,q_k}(z_0,\,z_1,% \ldots ,z_k-1)\ ,\ k+\sum_{i\neq k}\,z_i\neq 0\ \cdot \end{equation} The $q_k$-integration in $A_k^{\,q_k}$ can now be performed : $$ \int d^4q_k\,[(q_k+p_2)^2]^{z_k-1}\,[(q_{k-1}+q_k)^2]^{z_{k-1}}= $$ \begin{equation} \label{eq24}\pi ^2\frac{\Gamma (-1-z_k-z_{k-1})}{\Gamma (1-z_k)\,\Gamma (-z_{k-1})}\,{\rm B}(1+z_k,2+z_{k-1})\,[(q_{k-1}-p_2)^2]^{1+z_k+z_{k-1}}{\rm % ,} \end{equation} \noindent for $-1<Re(z_k)<1\,,\ -2<Re(z_{k-1})\,,\ Re(z_k)+Re(z_{k-1})<-1.$% We then obtain the reccurence formula $$ A_k(z_0,z_1,\ldots ,z_k)=\frac{\pi ^2}{\,[k+\sum_{j\neq k}z_j]}\,\frac{% \,\Gamma (-1-z_k-z_{k-1})}{\,\Gamma (-z_k)\,\Gamma (-z_{k-1})}\times $$ \begin{equation} \label{eq25}\,B(1+z_k,\,2+z_{k-1})\,A_{k-1}(z_0,z_1,\ldots ,z_{k-2},\,1+z_k+z_{k-1})\ , \end{equation} where $Re(z_k)<0$ and $A_k\quad (k=1,3,\ldots )$ is also defined by eq.(19) but with $q_{k+1}=-p_2$ . Note that eq.(25) obviously holds for $% k=1,3,5,\ldots $ too. Introducing the function \begin{equation} \label{eq26}D_k(z_0,z_1,\ldots ,z_k)=\prod_{\imath =0}^k\,\frac{\Gamma (-z_i) }{\Gamma (2+z_\imath )}\,A_k(z_0,z_1,\ldots ,z_k)\quad k=0,1,2,\ldots , \end{equation} the reccurence formula in eq.(25 ) takes the simple form \begin{equation} \label{eq27}D_k(z_0,z_1,\ldots ,z_k)=\frac{\pi ^2}{(1+z_k)(k+\sum_{\jmath \neq k}z_\jmath )}\,\,D_{k-1}(z_0,z_1,\ldots ,z_{k-2},1+z_{k-1}+z_k) \end{equation} and can be easily solved to give $$ D_k(z_0,z_1,\ldots ,z_k)=\pi ^{2k}\,\prod_{m=1}^k\,\frac 1{(m+z_0+\cdots +z_{m-1})(k+1-m+z_m+\cdots +z_k)} $$ \begin{equation} \label{eq28}\times \frac{\Gamma (-k-\sum_{\imath =0}^k\,z_\imath )}{\Gamma (2+k+\sum_{\imath =0}^k\,z_\imath )}\,s^{k+\sum_iz_\imath }. \end{equation} This formula holds provided that $$ -2<Re(z_m)\quad m=0,1,\ldots ,k-1\ ; $$ $$ -1<k-m+Re(z_m+\cdots +z_k)<0\quad m=1,2,\ldots ,k)\ ; $$ $$ k+Re(z_0+\cdots +z_k)<0\ ; $$ \begin{equation} \label{eq29}m+z_0+\cdots +z_{m-1}\neq 0\quad m=1,2,\ldots ,k \end{equation} as can be easily deduced from the restrictions which follow eqs.(19), (24) and (25). Substituting eq(28) in eq.(18) we obtain $$ I_\delta ^k(s)=\frac{4^k\pi ^{2(2k+1)}}{(2\pi \imath )^{k+1}}\,\prod_{\imath =0}^k\,\int_{c_\imath -\imath \infty }^{c_\imath +\imath \infty }\,dz_\imath \,(\delta ^2)^{2-n+z_\imath }\,\Gamma (n-2-z_i)\,\gamma ^{*}(n-2-z_i,-\frac{% \kappa _\imath }{\delta ^2})\times $$ \begin{equation} \label{eq30}\prod_{m=1}^k\frac 1{(m+z_0+\cdots +z_{m-1})(k+1-m+z_m+\cdots +z_k)}\times \end{equation} $$ \frac{\Gamma (-k-\sum_{\imath =0}^kz_\imath )}{\Gamma (2+k+\sum_{\imath =0}^kz_\imath )}\,\left( \frac s4\right) ^{k+\sum_\imath z_\imath } $$ with $-1<c_\imath <0\quad (\imath =0,1,\ldots ,k)$ and $k+c_0+\cdots +c_k<0$. The $z_\imath $- integrals can be evaluated by collapsing their contours to the right and using residue calculus. Since we are only interested in the imaginary part acquired by the amplitude when $s\rightarrow e^{-\imath \pi }s $ , we only keep contributions proportional to $\ln s$. The relevant contributions to the first $k\,$ $z_\imath $-integrals $(\imath =0,1,\ldots ,k-1)$ then come from the simple poles of the functions $\Gamma (n-2-z_i)\,(\imath =0,1,\ldots ,k-1)$ whereas the $z_k$-integral gets contributions from the double poles of the product $\,\Gamma (n-2-z_k)\Gamma (-k-\sum_{\imath =0}^kz_\imath )$ where the first $k\,\ z_\imath $'s $% (\imath =0,1,\ldots ,k-1)$ have already been substituted by integers. The final result is% $$ Im\,I_\delta ^k(se^{-\imath \pi })=4^k\pi ^{4k+3}\sum_{n-1\leq l_0,l_1,\ldots ,l_k}\,\prod_{\imath =0}^k\,\frac{(\kappa _\imath )^{l_\imath +1-n}}{(l_\imath +1-n)\,!}\times $$ \begin{equation} \label{eq31}\prod_{m=1}^k\,\frac 1{[l_0+\cdots +l_{m-1}][l_m+\cdots +l_k]}% \times \frac{\left( s/4\right) ^{\sum_{\imath =0}^k\,l_\imath -1}}{% [\sum_{\imath =0}^k\,l_\imath -1]\,!\,[\sum_{\imath =0}^k\,l_\imath ]\,!} \end{equation} and turns out to be again $\delta $-independent. Performing the $\rho _\imath ^2\,,\widetilde{\rho }_\imath ^2$ -integrals in eq.(16) we finally obtain the contribution of the ladder graph with $k+1$ rungs $% (k=0,2,4,\ldots )$ to $\sigma _{inc}$ :% $$ \sigma _{inc}^k=\frac 1{\pi \,s}\,\left[ \,4\pi ^4\,C\,(\pi ^2v^2)^{n-3-2\alpha }\right] ^{k+1}\,\,\sum_{n-1\leq l_0,l_1,\ldots ,l_k}\,\prod_{\imath =0}^k\frac{\left[ \Gamma (l_\imath +\alpha +2-n)\right] ^2}{\Gamma (l_\imath +2-n)}\,\times $$ \begin{equation} \label{eq32}\prod_{m=1}^k\,\frac 1{[l_0+\cdots +l_{m-1}][l_m+\cdots +l_k]}% \times \frac{\left( s/4\pi ^2v^2\right) ^{\sum_{\imath =0}^k\,l_\imath }}{% [\sum_{\imath =0}^k\,l_\imath -1]\,!\,[\sum_{\imath =0}^k\,l_\imath ]\,!} \end{equation} The multiple Series found for $\sigma _{inc}^k$, as it stands, looks very complicated to be handled. We shall attempt to get an estimate by finding suitable upper and lower bounds. In order to achieve this, we shall make extensive use of the inequalities% $$ l_0\,l_1\cdots l_k\leq \frac 1{(k+1)!}\,(l_0+l_1+\cdots +l_k)^{k+1}\quad , $$ \begin{equation} \label{eq33}\frac 1{l_0\,l_1\cdots l_k}>\frac{(k+1)!}{(l_0+l_1+\cdots +l_k)^{k+1}}\,\cdot \end{equation} Taking into account that $\alpha >2$ and defining $a=[\alpha ]+1$ we find that \begin{equation} \label{eq34}\sigma _l^k<\sigma _{inc}^k<\sigma _u^k\,, \end{equation} where $$ \sigma _l^k=\frac 1{\pi \,s}\,D^{k+1}\,2^k\,\,\,\sum_{n-1\leq l_0,l_1,\ldots ,l_k}\,\frac 1{\left[ \sum_{\imath =0}^kl_\imath \right] ^{n(k+1)-1}}\,\frac{% \prod_{\imath =0}^k\Gamma (l_\imath )}{\left[ \Gamma (\sum_{\imath =0}^kl_\imath )\right] ^2}\left[ \frac s{4\pi ^2v^2}\right] ^{\sum l_\imath }\,, $$ $$ \sigma _u^k=\frac 1{\pi \,s}\,D^{k+1}\frac 1{(n-1)^{2k}}\frac 1{(k+1)^{2a}}% \frac 1{(k!)^{2(a+1)}}\left( \frac{a^a}{a!}\right) ^{2(k+1)}\times $$ \begin{equation} \label{eq35}\sum_{n-1\leq l_0,l_1,\ldots ,l_k}\,\left[ \sum_{\imath =0}^kl_\imath \right] ^{2a(k+1)-1}\,\frac{\prod_{\imath =0}^k\Gamma (l_\imath )}{\left[ \Gamma (\sum_{\imath =0}^kl_\imath )\right] ^2}\left[ \frac s{4\pi ^2v^2}\right] ^{\sum l_\imath }, \end{equation} with $D=4\pi ^4\,C\,(\pi ^2v^2)^{n-3-2\alpha }\,.$ It is now clear that the next step must be the study of the multiple Series \begin{equation} \label{eq36}\Sigma _k(x)=\frac 1x\sum_{l_\imath =1}^\infty \frac{\Gamma (l_0)\cdots \Gamma (l_k)}{\left[ \Gamma (l_0+l_1+\cdots +l_k)\right] ^2}% \,x^{l_0+l_1+\cdots +l_k}\,,\quad x=\frac s{4\pi ^2v^2}\,\,\cdot \end{equation} Then, $\sigma _l^k$ and $\sigma _u^k$ could be recovered by integrating or differentiating $\Sigma _k(x)$ with respect to $\ln (x)$ a suitable number of times. The fact that the $l_\imath $-summations in the definition of $% \sigma _l^k$ and $\sigma _u^k$ start at $n-1$ cannot change the results in any fundamental way. The Laplace transform of $\Sigma _k(x)$ is% $$ S_k(t)=\int_0^\infty dx\,e^{-xt}\Sigma _k(x)= $$ \begin{equation} \label{eq37}\sum_{l_\imath =1}^\infty \frac{\Gamma (l_0)\,\Gamma (l_1)\cdots \Gamma (l_k)}{\Gamma (l_0+l_1+\cdots +l_k)}\,t^{-(l_0+l_1+\cdots +l_k)} \end{equation} We can now use the integral representation for the generalized beta function \begin{equation} \label{eq38}\frac{\Gamma (l_0)\,\Gamma (l_1)\cdots \Gamma (l_k)}{\Gamma (l_0+l_1+\cdots +l_k)}=\int_0^1\,\prod_{\imath =0}^k\,d\alpha _{\imath \;}\prod_{\jmath =0}^k\,\alpha _\jmath ^{l_\jmath -1}\,\delta (1- {\textstyle \sum }\alpha_\imath ) \end{equation} to get \begin{equation} \label{eq39}S_k(t)=\int_0^1\,\frac{\prod d\alpha _\imath }{\prod \alpha _\imath }\delta (1-{\textstyle \sum } \alpha _\imath )\,\sum_{l_\imath =1}^\infty \left[ \frac{\alpha _0}t\right] ^{l_0}\cdots \left[ \frac{\alpha _k}t\right] ^{l_k}. \end{equation} The summations are now decoupled and can be readily performed provided that $% \mid \alpha _\imath /t\mid <1.\quad $ We obtain \begin{equation} \label{eq40} S_k(t)=\int_0^1 \frac { \prod d\alpha _i\,\delta (1-{\textstyle \sum } \alpha _\imath ) } {(t-\alpha _0)\,(t-\alpha _1)\cdots (t-\alpha _k) } \quad \cdot \end{equation} The inverse Laplace transform of the expression above can be found and the answer is \begin{equation} \label{eq41} S_k(x)=\int_0^1 \prod d\alpha _i\,\delta (1-{\textstyle \sum } \alpha _\imath) \sum_{m=0}^k\,\frac {e^{\alpha _mx}}{P_m(\alpha _m)} \end{equation} where $P_m(y)$ is a polynomial given by \begin{equation} \label{eq42}P_m(y)=\,\frac{\prod_{\imath =0}^k(y-\alpha _\imath )}{y-\alpha _m}\quad \cdot \end{equation} Now the $\alpha _\imath -$ integrations can be performed and we end up with the following recursive formula \begin{equation} \label{eq43}\Sigma _k(x)=e^{\frac xk}\int_0^xdz\,e^{-\frac zk}\int_0^1d\rho \,\Sigma _{k-1}(\rho \,z)\,\quad \cdot \end{equation} This integral equation can be transformed into an integrodifferential equation \begin{equation} \label{eq44}\frac d{dx}\Sigma _k(x)=\frac 1k\,\Sigma _k(x)+\frac 1x% \,\int_0^x\,\Sigma _{k-1}(z)\,dz\,\, \end{equation} or a differential equation \begin{equation} \label{eq45}x\,\frac{d^2}{dx^2}\,f_k(x)\,+\,\left[ \left( 2-\frac 1k\right) x+1\right] \,\frac d{dx}f_k(x)\,+\,\left( 1-\frac 1k\right) \left( x+1\right) \,f_k(x)=\,f_{k-1}(x)\,, \end{equation} where \begin{equation} \label{eq46}f_k(x)=e^{-x}\,\Sigma _k(x)\,\cdot \end{equation} When $x$ is large, the differential equation reduces to \begin{equation} \label{eq47}\,\frac{d^2}{dx^2}\,f_k(x)\,+\,\left[ 2-\frac 1k\right] \,\frac d% {dx}f_k(x)\,+\,\left[ 1-\frac 1k\right] \,f_k(x)=\frac{\,f_{k-1}(x)}x\quad \cdot \end{equation} One particular integral of this equation is \begin{equation} \label{eq48}f_k(x)=\frac k{k-1}\,\frac 1{1+\frac d{dx}}\,\frac 1{1+\frac k{% k-1}\frac d{dx}}\,\frac{\,f_{k-1}(x)}x \end{equation} with $\,f_1(x)=1\,.$ \noindent It is now obvious that the leading order solution of eq.(45 ) can be written as \begin{equation} \label{eq49}f_k(x)=\frac k{x^{k-1}} \end{equation} leading to \begin{equation} \label{eq50}\Sigma _k(x)=k\ \frac{e^x}{x^{k-1}}\,\cdot \end{equation} We have already pointed out that the upper and lower bounds for $\sigma _{inc}^k$ which were defined in eq.( 35) can be related to derivatives or integrals of $\Sigma _k(x)$ with respect to$\ \ln x$. Such operations, however, cannot modify the form of $\Sigma _k(x)$ in an essential way since the exponential growth cannot be affected . We do expect a change in the leading power of $x$ and of course the constant coefficient will be different. We conclude that $\sigma _l^k\sim c_l^k\,e^x\,x^{-m_1},\,\sigma _u^k\sim c_u^k\,e^x\,x^{-m_2}$ where the constants $c_l^k$, $c_u^k$ , $m_1$ , $m_2$ can in principle be calculated. The cross section $\sigma _{inc}^k$ being bound between two exponentials, can only behave exponentially, possibly modified by an asymptotic Series of inverse powers of $s$ . Taking into account that $\sigma _{inc}^k$ for $k>1$ is highly supressed by the small $% D^{k+1}$ factor containing the instanton 't Hooft factor, we deduce that the contribution of all ladder graphs for $k>1$ is negligible and cannot alter the high energy behavior of the leading order result. In particular they do not affect the possible validity of the ZMS picture, that is based on the instanton-antiinstanton chain graphs only. This result is not totally unexpected. It is known that to each shaded blob of fig.2, which represents exchange of any number of bosons and $4n_g-2$ fermions between an instanton and an intiinstanton, corresponds an exponentially growing factor, while, to each instanton or antiinstanton, an exponentially small 't Hooft factor. Since the number of instanton or antiinstanton vertices outnumbers the number of the multiparticle-exchange blobs by a factor of two, we can expect that the contribution of such ladders is supressed compared to the leading semiclassical result. Moreover, the fact that not all the momentum flows through each rung makes at the end the ladder graphs grow only exponentially with $s$. This is in contrast to the case of a linear chain where all momentum flows between the instanton and the antiinstanton, creating thus an exponential growth that can counterbalance the suppression effect of the 't Hooft factors. We thank Q. Shafi and C. Bachas for collaborating in early stages of this work.\
\section{Introduction} The Canada-France Redshift Survey (CFRS) is aimed at the detection of faint galaxies among objects with 17.5$\le I_{AB} \le$ 22.5, with no discrimination against any other properties. The final catalog comprises 943 objects, of which 736 are extragalactic. The most common methods used to produce samples of quasars or active galactic nuclei (AGN) exploit their unusual colours, strong emission lines, variability, radio emission, X-ray emission, etc., and therefore are always biassed in some sense. Hewett and Foltz (1994) stress that if a completely unbiassed sample is not possible, an accurate assessment of the probability of detection as a function of absolute magnitude, redshift and spectral energy distribution must be made, and this is frequently lacking. Even though very effective and efficient survey techniques have been developed, there is always a danger that there may exist a subset of the population that might have been overlooked. Indeed, Webster(1995) has recently suggested that many more red quasars exist than previously supposed. Although the results of surveys based on techniques other than the popular UVX method can be examined to see if any such red population exists, surveys like CFRS can be considered as the ultimate check since spectroscopic observations are made for every object regardless of colour, morphology or spectroscopic properties. For example, the CFRS survey could detect red, radio-quiet, X-ray quiet quasars should they exist. On the other hand, since quasars constitute only $\sim$ 1\% of the surface density of objects on the sky at these magnitudes, CFRS cannot be expected to be an efficient survey for quasars. The fundamental limit to CFRS as a quasar survey is the signal-to-noise at which a broad emission line object could be identified. In this paper, the properties of the six objects which would commonly be classified as quasars in major surveys are compared with those from other samples. Although distinction is often made between quasars and other AGNs, particularly Seyfert galaxies, such a distinction is not usually made in the basic surveys: any point-like object with strong, broad emission lines is usually classified as a quasar. Subsequently, further subdivisions are made based on luminosities and whether or not the object is extended under typical ground-based seeing conditions. In this paper, we adopt the traditional survey definition, and simply refer to all six strong emission line objects as quasars, partly to avoid confusion with other galaxies in our survey which exhibit ``AGN activity'' in their nuclei (e.g., Tresse et al. 1994). Our CFRS quasar sample, although very small, is relatively unique and hence it is of interest to compare it with samples derived by other methods. Colless et al. (1991) reported a similar analysis based on detection of two quasars in a faint blue-selected sample, $21 < b_{J} < 23.5$. Throughout this paper it is assumed that $H_\circ=50$ km sec$^{-1}$ Mpc$^{-1}$ and $q_{\circ}=0.5$. \section{Observations} As part of the Canada-France Redshift Survey, spectra were obtained of over 1000 objects with 17.5$\le I_{AB} \le$ 22.5 in five different high galactic latitude fields selected to produce a fair sample of extragalactic objects (Lilly et al 1995a). The CFHT MOS spectrograph was used to cover the spectral region 4200 -- 8600\AA\ with a resolution of 7\AA\ per CCD pixel, and slits were used which yielded a spectral resolution of FWHM $\sim$35\AA. $BVIK$ photometry was also obtained for most objects in the spectroscopic sample (Lilly et al. 1995a). Details of all the observations, reduction techniques and sensitivity are presented by Le F\`evre et al. (1995a), Hammer et al. (1995), Lilly et al. (1995b) and Crampton et al. (1995). Eight one hour exposures were usually obtained of all targets so that spurious features such as cosmic rays were easily rejected, and features such as strong emission lines were readily apparent in all the individual spectra. The final CFRS catalog includes 591 galaxies with z $<$ 1.3, 200 stars, 146 unidentified objects and 6 quasars. Colour and morphological information indicates that most of the unidentified objects are likely to be galaxies (Crampton et al. 1995) but 7 objects are quite compact and have colours not very different from the quasars. The minimum equivalent width of emission lines that would have been detected in the spectra of these seven objects is $\sim$30\AA. Emission lines in the observed frame of at most a few percent of quasars at z $>$ 2 would have been missed, if the equivalent width distributions of Francis et al. (1992) and Hartwick and Schade (1990) are assumed. The Mg II equivalent width distribution indicates that only $\sim$10\% of low redshift (z $<$ 1) quasars might have been missed, so it is unlikely that any quasars reside among the unidentified objects. Thus, only 6 quasars were detected among a total of 736 extragalactic objects. The total effective area surveyed was 112 arcmin$^2$. Table 1 summarizes the properties of the quasars. The first column gives the CFRS number (the first two digits represent the right ascension of the field). The second and third columns give 2000 coordinates, followed by the isophotal I$_{AB}$ magnitude ($I_{AB} = I$ + 0.48), redshift, absolute magnitude, $(V-I)_{AB}$ colour, spectral index and rest-frame equivalent widths of strong emission lines. The errors of the equivalent widths are estimated to be $\pm10$\AA\ for most of the lines, and $\pm5\AA$ for the lines of CFRS14.0198. Absolute magnitudes in the rest-frame B band were computed assuming a power-law continuum slope, $f_\nu=\nu^\alpha$, with $\alpha$ determined for each object from its $(V-I)_{AB}$ colour corrected for emission line contamination. As indicated above, not all of these ``quasars'' lie above the canonical division between quasars and Seyfert galaxies at $M_B = -23$ for ($H_\circ=50$ km sec$^{-1}$ Mpc$^{-1}$). It is interesting to note that no very high redshift quasars were discovered. Since our targets were selected by their $I$ magnitudes, there is certainly no selection bias against high redshift quasars, and since Ly$\alpha$ would not disappear from our spectroscopic bandpass until z $>$ 6, identification of high redshift quasars would have been straightforward had they been present. Spectra of the six quasars are presented in Figure 1 with the strongest emission lines marked. Finding charts from our deep $I$ band images are given in Figure 2. At least three of the quasars may be in groups or large structures of galaxies, as noted below. \subsection{Notes on individual quasars} \noindent {\bf CFRS00.0207 z=1.352} is surrounded by many faint galaxies of similar magnitude; the five CFRS galaxies indicated in Figure 2 have isophotal $I_{AB} \sim23.4\pm0.3$mag. It is conceivable that this quasar is embedded in a cluster similar to those recently proposed by Hutchings et al. (1994) and Matthews et al. (1994) around other z $>$ 1 quasars. \noindent {\bf CFRS03.0106 z=2.070} is 2\farcs3 from a $\sim$1.3 mag fainter galaxy. \noindent {\bf CFRS03.0603 z=1.048} has a companion (615) only 16$^{\prime\prime}$ away which has identical redshift (1.048). In the larger CFRS field in this direction, there are three more galaxies with redshifts within $\Delta$z = 0.01 bringing the total to five. We have spectra of only one other object shown in the field; CFRS03.0602 is a star. The image of CFRS03.0603 appears to be slightly extended compared to the nearby stars, and the measured ``compactness parameter'', Q , (Le Fevre, et al. 1986) also indicates that it is resolved. Typically, CFRS point source objects have Q $\le$1.3 (Crampton et al. 1995), while Q = 1.55 for this quasar. Since the absolute magnitude of this quasar is M$= -23.0$, it is at the border of the canonical Seyfert -- quasar classification. Nevertheless, since z $\sim$1, it must reside in a bright host galaxy. \noindent {\bf CFRS14.0198 z=1.6034:} The bright object (163) is an M star. \noindent {\bf CFRS14.1303 z=0.9859} appears to be part of a very large structure (spread over our entire field corresponding to 6.5 $h_{50}^{-1}$ Mpc projected dimension for $q_\circ=0$, and greater than 900 km/s in redshift space) containing an estimated 30 bright galaxies at z = 0.985 (Le F\`evre et al. 1994). The nearest of these is CFRS14.1262, only 17$^{\prime\prime}$ away. CFRS14.1275 is at z = 0.763 and CFRS 14.1327 is at z = 0.932. Strong lines of Ne III and Ne V are present in the spectrum of CFRS14.1303 (Fig. 1). \noindent {\bf CFRS14.1567 z=0.4787} is slightly extended, indicating that the host galaxy is visible. Since M$= -22.0$, this quasar is more properly classified as a Seyfert. Only two of the nearby galaxies have redshifts measured; both 1525 and 1541 have z $\sim$0.74. Even though there are indications that at least three of the quasars may be located in groups of galaxies, a more rigorous analysis (Le F\`evre et al 1995b) shows that only the group around 14.1303 is statistically significant. \section{Discussion} \subsection{Surface density} The effective area of the CFRS survey is 0.031 deg$^{2}$ so that the observed 6 quasars translate to a surface density of 200$^{+120}_{-80}$ deg$^{-2}$; the highest surface density observed to date. The errors enclose a 1$\sigma$ confidence interval (Gehrels 1986). The survey magnitude limit of $I_{AB}\le$22.5 corresponds to $B=23$ for quasars with a power-law spectral energy distribution ($f_\nu\propto \nu^{\alpha}$) with $\alpha=-0.5$. For comparison, the previously-faintest quasar surveys reached a limiting magnitude of $B\sim$22. The Koo and Kron (1988) survey has a limiting magnitude of m$_{J}$=22.5 ( $B \sim$m$_{J}+0.1$), but spectroscopic followup has not yet been completed (Majewski, private communication). Zitelli et al. (1992, hereafter Z92) have surveyed 0.35 deg$^{2}$ to m$_{J}=22$ (and a smaller area to m$_{J}$=20.85). Their survey yields a quasar surface density at m$_{J}$=22 of 115$\pm$17 deg$^{-2}$ and, for z$<$ 2.2, a surface density of 86$\pm17$ deg$^{-2}$. Boyle, Jones, and Shanks (1991) surveyed an area of 0.85 deg$^2$ to a similar limiting magnitude. Although their survey suffers from incompleteness for z $>$ 2.2 due to the UVX selection procedure, the surface density for quasars with z$<$2.2, $68\pm9$ deg$^{-2}$, is in reasonable agreement with the Z92 result. In total, 118 new quasars were discovered in these two surveys. The Z92 result is also consistent with the surface density estimated in a review of all major surveys by Hartwick and Schade (1990). They derive a surface density of 160 quasars deg$^{-2}$ at $B$ = 22.5 and z $<$ 3.3 (or 129 at z $<$ 2.2) by applying a completeness correction to the work of Koo and Kron (1988). Figure 3 shows a comparison of all of these results for z $<$ 3.3. The surface density of quasars, log N $<$ B deg$^{-2}$, from Hartwick and Schade (1990) (circles) is combined with the newer results from Boyle et al. (1991) (open triangle) and Z92 (solid triangle). The faintest point (solid square) from the present survey lies on an extrapolation of the data from previous surveys, although the error bar is large. An extrapolation of the Z92 quasar surface density using their slope of 0.40 for the log $N$ versus $M$ relation gives an expected number of 300 deg$^{-2}$ at $B = 23$ for all redshifts, or 200 deg$^{-2}$ at z $<2.2$. As noted above, although all of our quasars are at z $<2.2$, we clearly have no bias against objects at much larger redshifts. Our observed surface density of 200$^{+120}_{-80}$ deg$^{-2}$ is at most a mild contradiction to the extrapolation of the Z92 surface density in the sense that the expected number of low-redshift (z $<2.2$) objects is observed, but there is some deficit of high-redshift objects. The faintest X-ray counts where optical identifications have been made are dominated by active galactic nuclei (Shanks et al 1991). Analysis of the deepest ROSAT counts yield surface density estimates of 413 X-ray sources deg$^{-2}$ to a limiting flux of $2.5 \times 10^{-15}$ erg cm$^{-2}$ s$^{-1}$ (Hasinger et al. 1993) and fluctuation analysis (Barcons et al. 1994) yields an estimate of 900 to 1800 discrete sources deg$^{-2}$ brighter than $7 \times 10^{-16}$ erg cm$^{-2}$ s$^{-1}$. These counts (in the 0.5-2 KeV) band which might be taken as estimates of the quasar surface density are well above our own estimate. \subsubsection{Comparison with Boyle (1991) model} The expected number of quasars in various redshift bins (e.g, $0 <$ z $<0.5$, $0.5<$ z $<1.0$, ...) can be computed from the model of Boyle et al. (1991). We find good agreement between the expectations derived from the evolving model and our observations. Over the range $0 < $z$ < 3$ , 4.98 quasars are predicted by the model compared to 6 observed. The redshift distribution of the observed quasars also agrees well with the model prediction, as shown in Figure 4. Agreement in luminosity is also good: 2.4 objects or $\sim$50\% should be AGN with $M_B>-23$, compared to the real sample which has 3, also 50\%. In summary, the redshift and luminosity distributions of the quasars in our sample agree very well with the predictions of the model of the evolving luminosity function of Boyle et al. (1991), demonstrating consistency of the model with observations in an $M-$z regime which represents a significant extrapolation beyond the regime where it was derived. \subsubsection{Relative numbers of quasars and Seyferts} In the preceding, we have made no distinction between luminous quasars ($M_B<-23$, $H_\circ=50$, $q_\circ=0.5$) and lower luminosity AGNs or Seyferts. An interesting point made by Z92 is that the counts of the Seyfert galaxies are rising more steeply (with a slope of $\sim0.7$ in the log $N$ vs $m$ relation) than the more luminous objects (whose slope is $\sim 0.4$). They note that the ratio of Seyfert to quasar counts increases from 0.1 at m$_{J}<20$ to $\sim 0.3$ at m$_{J}<21$ and is $\sim 0.5$ at m$_{J}<22$. In the faintest half-magnitude bin the ratio reaches unity. In our very faint sample, the ratio of Seyfert to quasar counts also equals 1, confirming the trend noted by Z92. \subsection{Continuum properties} The continua of active galactic nuclei (AGN) are frequently characterized as power-law $f_\nu \propto \nu^\alpha$ spectra, although this is a simplification when the continuum is considered over a large range in wavelength (e.g., Neugebauer et al. 1979). For example, the continuum shape in the optical-ultraviolet region is dominated by the ``blue-bump'' feature which may be the signature of a hot accretion disk (Elvis et al. 1994). Since many of the quasars previously studied have been selected because they have blue continua (e.g., in ultraviolet-excess surveys), there may be biases in the statistics related to continuum shape. Hence, it is of interest to examine the continuum slopes in our sample where no such bias can exist. The optical-UV region of quasar spectra are crowded with emission lines, Balmer continuum emission, and a number of FeII emssion features (e.g., Francis et al. 1991, hereafter F91). This makes the definition of the continuum difficult, particularly when only a restricted wavelength range (and one which varies with redshift) is available. In general, our spectra were well-fitted by power laws, but estimates of $\alpha$ made in this way are subject to considerable error, due to errors in the centering of any individual object among the $\sim$80 objects in our multi-slit masks. Consequently, the continuum spectral index was estimated from our $V-I$ colours (corrected for the contribution of emission lines), assuming a power law. The slopes, listed in Table 1, range from $\alpha=0.2$ to $\alpha=-2.0$ with a mean of $-1.0$. The error is estimated to be $\sim \pm 0.2$ in $\alpha$ One of the largest and most complete surveys with which to compare our results is the Large Bright Quasar Survey (LBQS). F91 give the distribution of power-law slopes measured in the wavelength range 1450 -- 5050\AA\ for 688 LBQS (m $\le18.85$) quasars. They find a mean $\alpha = - 0.32$. Francis et al. (1992) found a higher value using a principal component analysis method on a smaller subsample of the LBQS, but since the rest-frame continuum window used was very blue, the earlier analysis is better matched to our data. Interestingly, the results from our small sample span nearly the entire range in continuum slope evident in the much larger sample of F91. Half of our spectral indices are redder than $-1$ whereas at most 10\% of those in F91 are as red as this. Furthermore, 5/6 of our spectral indices are redder than the median of the sample of F91. However, the 95\% confidence interval (Gehrels 1986) for the ratio $R$ of red-to-total objects derived from our observations is $0.36 < R < 0.996$, so that this result is not highly significant. There is no reason to believe that the survey technique employed in the LBQS would have missed the reddest (or any) quasars seen here. On the contrary, Francis et al. (1992) claim that any quasar with a continuum bluer than $\alpha\sim -2$ and/or emission lines would have been detected. Thus, although there is some indication that our sample is rather red, it does not constitute a population that would have escaped detection in the LBQS or other surveys. It should be remembered, however, that there are substantial differences in the luminosity and redshift distributions between our sample and the LBQS. \subsection{Emission-line strengths} The rest-frame equivalent widths of the three principal emission lines common to most of our spectra are listed Table 1. The equivalent widths are consistent with the means given by F91 and those compiled by Hartwick and Schade (1990), indicating once again that the quasars in our sample are not unusual. Since there has always been a lingering concern that the average emission line strengths of quasars have been biased by their detection methods, this is reassuring. \section{Summary} The main goal of this paper was to compare the properties of this very faint (albeit small) sample of quasars with the properties of previously known samples. This sample was selected without any of the biases inherent in the most common survey techniques and therefore could potentially yield new insights into the population. The surface density of AGN at $B=23$ is $200^{+120}_{-80}$ deg$^{-2}$, the highest observed to date. This is consistent with an extrapolation of the faint counts due to Z92 and Boyle et al. (1991), and also with the lower limit found by Colless et al. (1991) based on a similar method to ours. We find that the space density of these quasars as a function of redshift and luminosity agrees very well with the predictions of the successful evolutionary model of Boyle (1991), even though our sample is deeper than any of the samples from which the model is derived. The colours of these faint AGN are within the range defined by previously known samples and would not have escaped detection in major surveys. We find an apparent surplus of red objects compared with other studies, but the numbers are small and the significance is low. The strengths of the emission lines are entirely consistent with those found in other surveys. In summary, we find no remarkable differences between any of the properties of these faint objects and the distributions of the same properties derived from previous surveys. All our quasars would have been found by previous search techniques. One of the quasars is in a structure consisting of at least 30 galaxies at z $\sim0.985$, and there is evidence that two others may be associated with groups of galaxies at a similar redshift.
\section{Introduction} \label{intro} Let $L$ be a perfect field of characteristic $p$. Recall that a polynomial $f(x) \in L[x]$ is called {\em additive} if $f(x+y)=f(x)+f(y)$ identically. It is easy to see that a polynomial is additive if and only if it is of the form $$f(x)=a_0 x + a_1 x^p + a_2 x^{p^2} + \cdots + a_n x^{p^n}.$$ The set of additive polynomials forms a noncommutative ring in which $(f \circ g)(x)=f(g(x))$. This ring is generated by the scalar multiplications $x \mapsto ax$ for $a \in L$, and by $\tau(x)=x^p$. In this ring we can write the $f$ above as $a_0 + a_1 \tau + a_2 \tau^2 + \cdots + a_n \tau^n$, and we will denote the ring by $L\{\tau\}$, using braces instead of brackets to remind ourselves that it is a {\em twisted} polynomial ring in $\tau$. The indeterminate $\tau$ does not commute with elements of $L$ (which are acting as scalar multiplications); instead if $a \in L$ then $\tau a = a^p \tau$ as additive polynomials. Following Ore~\cite{Or1}, we define the {\em adjoint} of $f$ to be the expression $$f^\ast(x) = a_0 x + a_1^{1/p} x^{1/p} + a_2^{1/p^2} x^{1/p^2} + \cdots + a_n^{1/p^n} x^{1/p^n}.$$ This defines a function on $L$ (and its algebraic closure) since $L$ is perfect. Goss observed that the kernel of $f$ (the set of zeros of $f$ in a fixed algebraic closure of $L$) generates the same field extension as the kernel of $f^\ast$, by using a formula of Ore~\cite{Or1} which expresses the zeros of $f^\ast$ explicitly in terms of determinants involving the zeros of $f$. One might ask whether there is a natural Galois-equivariant isomorphism between $\ker f$ and $\ker f^\ast$. In general, the answer is no. But, as was discovered independently by the author and N. Elkies~\cite{elkies}, there is something which is just as good: \begin{unnumbered} There exists a natural Galois-equivariant pairing $$\ker f \times \ker f^\ast \rightarrow {{\Bbb F}_p}.$$ \end{unnumbered} One of the main results of this paper is the generalization of this theorem to the situation where $f$ is an additive {\em power series}. Let $C$ be an algebraically closed field of characteristic $p$ which is complete with respect to a (non-archimedean) absolute value $|\;|$. If $f \in C[[z]]$ is an additive power series, its adjoint is a ``linear fractional power series'' of the form $$g(z) = b_0 z + b_1 z^{1/p} + b_2 z^{1/p^2} + \cdots$$ It is easy to see that such a series converges for all $z$ when the $b_i$ tend to zero. In this case, we will show that $g$ defines a continuous group homomorphism $C \rightarrow C$, which is surjective and open if nonzero. One can see immediately that if $b_0 \not = 0$ then $g(z) = b_0 z + \text{(smaller terms)} \not= 0$ for large $z$, so $\ker g$ is a closed and bounded subgroup of $C$. In fact, we prove that $\ker g$ is compact, and if infinitely many $b_i$ are nonzero, it is isomorphic to ${{\Bbb F}^\omega_p}$ as a topological group. Conversely, if $G$ is any compact subgroup of $C$, then $G$ is the kernel of such a $g$ if and only if the integral $\int_G v$ of the real-valued function $v(z)=-\log|z|$ with respect to Haar measure on $G$ equals $+\infty$. Also, $G$ essentially determines $g$. We next develop the theory of Newton polygons for such series. Finally we prove the analogue of our pairing theorem above, namely that for any everywhere convergent additive power series $f \in C[[z]]$, there is a natural pairing $$\ker f \times \ker f^\ast \rightarrow {{\Bbb F}_p}$$ which exhibits the compact group $\ker f^\ast$ as the Pontryagin dual of the discrete group $\ker f$. Actually we will generalize further by considering ${{\Bbb F}_q}$-linear power series for any power $q$ of $p$, and more substantially by considering {\em bi-infinite series} $$f(z) = \cdots + a_{-2} z^{1/q^2} + a_{-1} z^{1/q} + a_0 z + a_1 z^q + a_2 z^{q^2} + \cdots.$$ The kernels of these series are locally compact. In fact, many of our results extend to non-linear series for which the set of allowable exponents is $$E=\{mp^n : m,n \in {\Bbb Z}, m \ge 0 \}.$$ We will develop results in this general context, specializing when appropriate. Finally we give some applications of our results to the theory of Drinfeld modules. We construct a pairing for Drinfeld modules which behaves in many ways like the Weil pairing on abelian varieties. In addition, our results let us describe the topological module structure of the kernel of the adjoint exponential function of a Drinfeld module. We conclude with a few unanswered questions. \section{The rings ${\cal P}$ and ${\cal F}$ of fractional power series} \label{tworings} Let $C$ be an algebraically closed field containing ${{\Bbb F}_q}$ which is complete with respect to a (non-archimedean) absolute value $|\;|$. Let $V = \{ x \in C : |x| \le 1 \}$ be the valuation ring of $C$, let ${\frak m} = \{ x \in C : |x|<1 \}$ be its maximal ideal, and let $k=V/{\frak m}$ be the residue field. We will consider ``fractional power series'' in which the exponents belong to $$E=\{mp^n : m,n \in {\Bbb Z}, m \ge 0 \}.$$ Since we have unique $p$-th roots in $C$, for any $z \in C$ and $e = mp^n \in E$ we can interpret $z^e$ as $(z^{1/p^n})^m$. Define ${\cal P}$ to be the set of series $\sum_{e \in E} a_e z^e$ with coefficients $a_e$ in $C$, which converge at all $z \in C$. Since $E$ is countable, it is clear what convergence means: for each $z \in C$ and $r>0$, there must be only finitely many terms in the series of absolute value greater than $r$. And since we are in the non-archimedean situation, convergence at some $z$ of absolute value $r$ implies uniform convergence on all of $\Delta_r$. Furthermore, we define the set ${\cal F}$ of ``${{\Bbb F}_q}$-linear fractional power series'' as the set of bi-infinite series $$f(z) = \cdots + a_{-2} z^{1/q^2} + a_{-1} z^{1/q} + a_0 z + a_1 z^q + a_2 z^{q^2} + \cdots$$ which converge for all $z \in C$. Hence ${\cal F}$ is a subset of ${\cal P}$. For $f \in {\cal F}$ we will also write $$f = \cdots + a_{-2} \tau^{-2} + a_{-1} \tau^{-1} + a_0 + a_1 \tau + a_2 \tau^2 + \cdots,$$ thinking of $\tau$ as the operator $\tau(z)=z^q$ on $C$, and thinking of $a \in C$ acting as the scalar multiplication by $a$. Note that for $a \in C$, $\tau a = a^q \tau$ and $\tau^{-1} a = a^{1/q} \tau^{-1}$, as maps on $C$. We also define ${\cal F}^+$ as the set of $f \in {\cal F}$ such that $a_n=0$ for $n<0$ (these are just the convergent ${{\Bbb F}_q}$-linear power series), and ${\cal F}^-$ as the set of $f \in {\cal F}$ such that $a_n=0$ for $n>0$. There is a simple criterion for the convergence of linear series: \begin{prop} \label{convergence} Suppose $a_i \in C$ for $i \in {\Bbb Z}$. The bi-infinite series $$f(z) = \cdots + a_{-2} z^{1/q^2} + a_{-1} z^{1/q} + a_0 z + a_1 z^q + a_2 z^{q^2} + \cdots$$ converges for all $z \in C$ if and only if the following two conditions hold: \begin{enumerate} \item $\lim_{n \rightarrow -\infty} |a_n| = 0$. \item $\lim_{n \rightarrow +\infty} |a_n|^{1/q^n} = 0$. \end{enumerate} \end{prop} \begin{pf} By the well-known formula for the radius of convergence of a power series, the right end of the series converges everywhere when $\lim_{n \rightarrow +\infty} |a_n|^{1/q^n} = 0$. The left end converges at $z$ when $\lim_{n \rightarrow -\infty} |a_n z^{q^n}| = 0$. But for $z$ nonzero, this is equivalent to $\lim_{n \rightarrow -\infty} |a_n| = 0$, since $\lim_{n \rightarrow -\infty} |z^{q^n}| = 1$. \end{pf} For $f,g \in {\cal P}$, we define $f+g$ and $f \cdot g$ in the obvious way, by expanding and grouping terms with the same exponent. \begin{prop} $({\cal P},+,\cdot)$ is a commutative ring. \end{prop} \begin{pf} The only nontrivial part is to check that the sum and product are actually everywhere convergent. In fact, this is easy as well, since the sum or product of two convergent series is a doubly infinite convergent series, even before grouping terms. \end{pf} If $f,g \in {\cal F}$, we can still use the definitions of addition and multiplication above, but $f \cdot g$ need not be an element of ${\cal F}$, so ${\cal F}$ is not a subring of ${\cal P}$. Nevertheless, we will make ${\cal F}$ into a ring by using composition in place of multiplication. In fact, we will define the composition $f \circ g$ of two general elements $f(z)=\sum_{d \in E} a_d z^d$ and $g(z)=\sum_{e \in E} b_e z^e$ of ${\cal P}$. First, for integral $m \ge 0$, let $g^m$ denote $g$ multiplied by itself $m$ times. For $n \in {\Bbb Z}$, define $g^{p^n} = \sum_{e \in E} b_e^{p^n} z^{ep^n}$, which is consistent with the previous sentence. For $d=mp^n \in E$, define $g^d = \left( g^m \right)^{p^n}.$ Finally define $$f \circ g = \sum_{d \in E} a_d g^d,$$ in which we expand and group terms of the same exponent. If $f=\sum_{i \in {\Bbb Z}} a_i \tau^i$ and $g=\sum_{j \in {\Bbb Z}} b_j \tau^j$ are elements of ${\cal F}$, then the above definition simplifies to $f \circ g =\sum_{n \in {\Bbb Z}} c_n \tau^n$ where $$c_n = \sum_{i+j=n} a_i b_j^{q^i}.$$ \begin{prop} \label{composition} For $f,g \in {\cal P}$, $f \circ g$ converges to an element of ${\cal P}$, and represents the composition of the maps $f$ and $g$. \end{prop} \begin{pf} Fix $r>0$. Since the series $g$ converges uniformly on $\Delta_r$, we may let $R = \sup |g(\Delta_r)|$. Fix $z$ of absolute value $r$. Then from the definition of $g^d$ it is clear that all terms in the expansion of $g^d$ (for fixed $d \in E$) are bounded in absolute value by $R^d$, with only finitely terms of absolute value greater than $s$, for each $s>0$. Fix $t>0$. There are only finitely many $d$ such that $|a_d|R^d > t$ (since $f$ is convergent), so there are only finitely many $d$ such that $a_d g^d$ has terms of absolute value greater than $t$, and the previous sentence implies that even for such $d$, there are at most finitely many such terms. Hence even before grouping terms, the series for $f \circ g$ converges at $z$, and this implies that the grouping gives a convergent series for each coefficient of $f \circ g$ as well. \end{pf} \begin{prop} \label{ring} $({\cal F},+,\circ)$ is a noncommutative ring containing $C$. The center of ${\cal F}$ is ${{\Bbb F}_q}$. \end{prop} \begin{pf} By the previous proposition and the explicit form of the definition for linear series, if $f,g \in {\cal F}$, then $f \circ g \in {\cal F}$. Hence the first part is clear. The center of ${\cal F}$ clearly contains ${{\Bbb F}_q}$. On the other hand, if $c \in C$ is transcendental, direct computation shows that its centralizer in ${\cal F}$ is $C$, and the only elements of $C$ which commute with $\tau$ are those in ${{\Bbb F}_q}$. \end{pf} \begin{prop} \label{uniformcontinuity} Any $f \in {\cal P}$ defines a continuous function $f: C \rightarrow C$. The continuity is uniform on each bounded subset of $C$. \end{prop} \begin{pf} It suffices to show $f$ is uniformly continuous on $\Delta_R$ for each $R>0$. Given $\epsilon >0$, there are only finitely many terms $a_e z^e$ in $f$ for which $|a_e| R^e >\epsilon$, since $f$ is everywhere convergent. The sum of these is clearly uniformly convergent on $\Delta_R$, and the sum of the rest of the terms is bounded in absolute value by $\epsilon$ for all $z \in \Delta_R$, so $f$ is uniformly continuous on $\Delta_R$ as well. \end{pf} \begin{prop} \label{nonvanishing} If $f \in {\cal P}$ is nonzero as a series, then the function it defines on $C$ is not identically zero. \end{prop} \begin{pf} Since $f(1)$ converges, there are only finitely many coefficients of $f$ larger than a given size, and by scaling by an element of $C$, we may assume the largest coefficients have absolute value 1. Then the reduction of $f$ modulo ${\frak m}$ is a fractional {\em polynomial} over the residue field $k$, i.e., the $p^n$-th root of a polynomial. Since $k$ is infinite, this reduction is nonzero at some $\bar{\alpha} \in k$. If $\alpha \in C$ is any lift of $\bar{\alpha}$, then $|f(\alpha)|=1$, so in particular $f(\alpha) \not= 0$. \end{pf} \begin{prop} \label{additive} If $f \in {\cal P}$ considered as a map on $C$ is ${{\Bbb F}_q}$-linear, then $f \in {\cal F}$. \end{prop} \begin{pf} Write $f= \sum_{e \in E} a_e z^e$. For each $c \in C$, the function $f(cz+z)-f(cz)-f(z)$ in ${\cal P}$ is identically zero, so by Proposition~\ref{nonvanishing} it is zero as a series. Hence for each $e=mp^n$, the coefficient of $z^e$ in it is zero. This coefficient is $a_e \left[ (c+1)^e - c^e - 1 \right]$. If $e$ is not a power of $p$ and $a_e \not=0$, then the $p^n$-th power of this coefficient will be a nonzero polynomial in $c$, and hence will not vanish for some value of $c$. This contradiction shows that $a_e=0$ for $e$ not a power of $p$. A similar argument with $f(cz)-cf(z)$ for $c \in {{\Bbb F}_q}$ shows that $a_e=0$ for $e$ not a power of $q$; i.e., $f \in {\cal F}$. \end{pf} \section{The topology and norm on ${\cal P}$ and ${\cal F}$} \label{topologynorm} We give ${\cal P}$ the ``bounded-open topology.'' For real $r>0$, define \begin{eqnarray*} \Delta_r & = & \{\, x \in C : |x| \le r \,\}, \\ \Delta_r' & = & \{\, x \in C : |x| < r \,\}. \end{eqnarray*} If $r,s>0$ let $${\cal P}(r,s)=\{\,f \in {\cal P} | f(\Delta_r) \subset \Delta_s \,\}.$$ Take these subsets of ${\cal P}$ as a subbasis for the neighborhoods of zero. For each $\alpha,\beta \in C$ and $c \in {{\Bbb F}_q}$, the set of $f \in {\cal P}$ such that $f(\alpha+\beta)=f(\alpha)+f(\beta)$ and the set of $f \in {\cal P}$ such that $f(c \alpha) = c f(\alpha)$ are closed subsets of ${\cal P}$. Hence by Proposition~\ref{additive}, ${\cal F}$ is a closed subset of ${\cal P}$. We give ${\cal F}$ the subspace topology, which is also the bounded-open topology on ${\cal F}$. For $f = \sum_{e \in E} a_e z^e \in {\cal P}$, define $$\|f\| = \sup\; \{\, |a_e|:e \le 1 \,\} \cup \{\, |a_e|^{1/e}:e \ge 1 \,\}.$$ The convergence of $f$ implies that this is a finite real number, and that the supremum is attained. If $f = \sum_{n \in {\Bbb Z}} a_n \tau^n \in {\cal F}$ this becomes $$\|f\| = \sup\; \{\, |a_n|:n \le 0 \,\} \cup \{\, |a_n|^{q^{-n}}:n \ge 0 \,\}.$$ This norm satisfies an ultrametric triangle inequality: $\|f+g\| \le \max\{\|f\|,\|g\|\}$ for any $f,g \in {\cal P}$. In showing that the norm is consistent with the bounded-open topology on ${\cal P}$ we will use the following. \begin{lemma} \label{termwise} Let $f = \sum_{e \in E} a_e z^e \in {\cal P}$, and let $r,s>0$. Then the following are equivalent: \begin{enumerate} \item $f \in {\cal P}(r,s)$. \item $a_e z^e \in {\cal P}(r,s)$ for all $e \in E$. \item $|a_e| r^e \le s$ for all $e \in E$. \end{enumerate} \end{lemma} \begin{pf} It is clear that~(2) and~(3) are equivalent, and that they imply~(1). In showing~(1) implies~(3), we can assume $r$ is the absolute value of some element of $C$, because if not, it is an increasing limit of such $r$, and the implication follows if it is known for those $r$. Then we can reduce to the case $r=1$ by composing $f$ with a scalar multiplication on the right. Now what we must show is that if $f$ is a nonzero element of ${\cal P}(1,s)$, a largest coefficient $b$ of $f$ is at most $s$ in absolute value. The mod ${\frak m}$ reduction of $b^{-1} f$ is the $p^n$-th root of a nonzero {\em polynomial}, so we can pick $x \in V$ which does not reduce to one of its roots, and find that $|b^{-1} f(x)|=1$. Hence $|b| = |f(x)| \le s$, since $f \in {\cal P}(1,s)$. \end{pf} \begin{prop} \label{agree} The bounded-open topology on ${\cal P}$ is the same as that induced by the norm $\|\;\|$. \end{prop} \begin{pf} It is clear from the definitions that if $0<s<1<r$, $$\|f\| \le s/r \implies f \in {\cal P}(r,s).$$ On the other hand, by Lemma~\ref{termwise}, if $0<s<1$, $$f \in {\cal P}(1,s) \cap {\cal P}(1/s,1) \implies \|f\| \le s.$$ \end{pf} \begin{theorem} \label{topologicalring} $({\cal P},+,\cdot)$ is a complete topological ring, and the composition map ${\cal P} \times {\cal P} \rightarrow {\cal P}$ is continuous. Also, $({\cal F},+,\circ)$ is a complete topological ring. \end{theorem} \begin{pf} The triangle inequality for $\|\;\|$ implies that addition is continuous. To check that multiplcation is continuous, we must show that if $f,g \in {\cal P}$, and $r,s>0$, then for $f_0,g_0$ sufficiently small, $$(f+f_0)\cdot(g+g_0)-f\cdot g \in {\cal P}(r,s).$$ Since $f$ and $g$ are bounded on $\Delta_R$, this follows for $f_0,g_0 \in {\cal P}(r,\epsilon)$ for sufficiently small $\epsilon>0$. Now let us show that ${\cal P}$ is complete. It suffices to show that if $f_1,f_2,\ldots$ is a sequence in ${\cal P}$ tending to zero, then $\sum_{i=1}^\infty f_i$ converges to another element of ${\cal P}$. For each $e \in E$, the coefficients of $z^e$ in $f_i$ tend to zero by Lemma~\ref{termwise}, so they sum to some $a_e \in C$. Interchanging the order of summation shows that $\sum_{i=1}^\infty f_i$ converges to $f \stackrel{\text{def}}{=} \sum_{e \in E} a_e z^e$ uniformly on bounded subsets, and that $f \in {\cal P}$. For the continuity of composition, we must show that for $f_0,g_0$ sufficiently small, \begin{equation} \label{compcont} (f+f_0) \circ (g+g_0) - f \circ g \in {\cal P}(r,s). \end{equation} Let $R=\sup |g(\Delta_r)|$. If $g_0 \in {\cal P}(r,R)$ and $f_0 \in {\cal P}(R,s)$, then \begin{equation} \label{firsthalf} f_0 \circ (g+g_0) \in {\cal P}(r,s). \end{equation} By Proposition~\ref{uniformcontinuity}, $f$ is uniformly continuous on $\Delta_R$, so if $g_0 \in {\cal P}(r,\epsilon)$ for sufficiently small $\epsilon$, \begin{equation} \label{secondhalf} f \circ (g+g_0) - f \circ g \in {\cal P}(r,s). \end{equation} Adding~(\ref{firsthalf}) and~(\ref{secondhalf}) yields~(\ref{compcont}). By restriction from ${\cal P}$ to ${\cal F}$, the ring operations on ${\cal F}$ are continuous. As remarked earlier, ${\cal F}$ is a closed subset of ${\cal P}$, so the completeness of ${\cal F}$ follows from that of ${\cal P}$. \end{pf} Although the ring operations for both ${\cal P}$ and ${\cal F}$ are continuous with respect the topology induced by $\|\;\|$, it is not true that $\|f \cdot g\| \le \|f\| \cdot \|g\|$ for all $f,g \in {\cal P}$; nor is it true that $\|f \circ g\| \le \|f\| \cdot \|g\|$ for all $f,g \in {\cal F}$. In fact, these can fail even if $f \in C$. The following proposition shows that this defect of $\|\;\|$ is unavoidable. (This is different from the theory of Banach algebras, in which if one has a norm for which multiplication is continuous in each variable, one can define a new norm with the sub-multiplicative property. See Section~3.1 of~\cite{kadison}.) \begin{prop} There is no norm $\|\;\|'$ on ${\cal P}$ inducing the same topology as above and satisfying \begin{enumerate} \item $\|f\|' \ge 0$ with equality if and only if $f=0$. \item $\|f \pm g\|' \le \|f\|' + \|g\|'$ \item $\|1\|' = 1$. \item $\|f \cdot g\|' \le \|f\|' \cdot \|g\|'$. \end{enumerate} Similarly, there is no norm $\|\;\|'$ on ${\cal F}$ inducing the same topology and satisfying (1), (2), (3), and (4) with $\cdot$ replaced by $\circ$. \end{prop} \begin{pf} Suppose there were such a norm on ${\cal P}$. As $\epsilon \rightarrow 0$ in $C$, $\epsilon z \rightarrow 0$ in ${\cal P}$, so $\|\epsilon z\|'<1$ for some nonzero $\epsilon \in C$. Then the series $1 + \epsilon z + (\epsilon z)^2 + \cdots$ should converge in ${\cal P}$ to an inverse of $1-\epsilon z$. This is impossible since $1-\epsilon z$ is zero at $z=1/\epsilon$. Similarly, since $1-\epsilon \tau$ has no inverse (with respect to composition), there cannot be a sub-multiplicative norm on ${\cal F}$. \end{pf} \section{A Class of Locally Compact Subspaces of $C$} \label{subspaces} For $z \in C^\ast$, let $$v(z)=-\log |z| \in {\Bbb R}$$ be the (additive) valuation associated with $|\;|$. By convention, $v(0)=+\infty$. If $f \in {\cal F}^+$, then as is well-known, $\ker f$ is a discrete sub-${{\Bbb F}_q}$-vector space of $C$, which is ``concentrated near infinity'' in the sense that if it is infinite, it is countable and its elements tend to infinity. If $f \in {\cal F}^-$, it will turn out that $G \stackrel{\text{def}}{=} \ker f$ is a {\em compact} sub-${{\Bbb F}_q}$-vector space of $C$, this time ``concentrated near zero'' in the sense that the integral $\int_G v$ of the valuation with respect to Haar measure on $G$ is $+\infty$. The kernel of a general (i.e., possibly infinite in both directions) element of ${\cal F}$ will belong to a certain hybrid of these two classes of vector spaces, which we now define. \begin{defn} Fix a real number $r>0$. Let ${\cal G}$ be the collection of sub-${{\Bbb F}_q}$-vector spaces $G$ of $C$ satisfying the following conditions: \begin{enumerate} \item $G \cap \Delta_r$ is compact. \item $\int_{G \cap \Delta_r} v = +\infty$. The integral is with respect to (any) Haar measure on $G \cap \Delta_r$. \item $G \cap \Delta_r$ has finite or countable index in $G$, and in the latter case, if $g_1,g_2,\ldots$ is a system of coset representatives, then $\lim_{n \rightarrow \infty} |g_n| = +\infty$. \end{enumerate} \end{defn} \begin{defn} Let $G$ be a sub-${{\Bbb F}_q}$-vector space of $C$. Let $\{\lambda_n\}_{n \in {\Bbb Z}}$ be a bi-infinite collection of elements of $C$, possibly terminating on either side. We say $\{\lambda_n\}_{n \in {\Bbb Z}}$ is a {\em descending basis} for $G$ if \begin{enumerate} \item The $\lambda_n$ are independent over ${{\Bbb F}_q}$ and the closure of their span is $G$. \item For each $m \in {\Bbb Z}$ (for which $\lambda_m$ exists), $\lambda_m$ is a smallest nonzero element of the ${{\Bbb F}_q}$-vector space spanned by $\{\, \lambda_n \mid n \le m \,\}$. \item The sequence terminates on the left, or $\lim_{n \rightarrow -\infty} |\lambda_n| = \infty$. \item The sequence terminates on the right, or $\sum_{n=1}^\infty v(\lambda_n) q^{-n} = +\infty$. \end{enumerate} \end{defn} \begin{prop} \label{class} Let $G$ be a sub-${{\Bbb F}_q}$-vector space of $C$. Then \begin{enumerate} \item The definition of ${\cal G}$ does not depend on the choice of $r>0$. \item $G$ has a descending basis if and only if $G \in {\cal G}$. \item If $G \in {\cal G}$, then $G$ is locally compact. \item For $G$ discrete, $G \in {\cal G}$ if and only if it is finite or countable, with its elements tending to infinity in the latter case. \item For $G$ compact, $G \in {\cal G}$ if and only if $\int_G v = +\infty$. \item If $G \in {\cal G}$, then any closed sub-${{\Bbb F}_q}$-vector space $H$ of $G$ also belongs to ${\cal G}$. \end{enumerate} \end{prop} \begin{pf} $\left.\right.$ \noindent {\em Proof of (2): } Suppose $G$ has a descending basis $\{\lambda_n\}$. Then the $\lambda_n$ are decreasing in size by condition~(2) in the definition, and $G \cap \Delta_r$ is the closure of the span of the $\{\lambda_n : n \ge n_0\}$ where $\lambda_{n_0}$ is the first $\lambda$ of absolute value at most $r$. If the sequence of $\lambda_n$ after $\lambda_{n_0}$ terminates, then $G \cap \Delta_r$ is finite, hence compact. Otherwise, these $\lambda_n$ form a sequence tending to zero by condition~(4), and we get a topological isomorphism ${{\Bbb F}^\omega_q} \rightarrow G \cap \Delta_r$ which sends $(a_0,a_1,a_2,\ldots)$ to $\sum_{i=0}^\infty a_i \lambda_{n_0+i}$. (This, together with Theorem~\ref{locallycompact}, proves the remark in the introduction regarding the topological group structure of the kernel of an element of ${\cal F}^-$.) Hence $G \cap \Delta_r$ is compact in any case. Thus condition~(1) of the definition of ${\cal G}$ is verified for $G$. Assume that the Haar measure on $G$ is normalized so that the open subspace $G \cap \Delta_r$ has measure 1. If $G \cap \Delta_r$ is finite, then $\int_{G \cap \Delta_r} v = +\infty$ since $0$ has positive measure. Otherwise, under the isomorphism above, the set of $(a_0,a_1,a_2,\ldots) \in {{\Bbb F}^\omega_q}$ such that the first nonzero $a$ is $a_i$ corresponds to a subset of $G \cap \Delta_r$ of measure $q^{-i}-q^{-i-1}$, and the function $v$ takes the value $v(\lambda_{n_0+i})$ everywhere on this subset, since otherwise some nontrivial combination of the $\lambda$'s of the same size as $\lambda_{n_0+i}$ would be closer to zero than $\lambda_{n_0+i}$, violating condition~(2) in the definition of decreasing basis. These sets cover ${{\Bbb F}^\omega_q}$ (except for the point 0), so $$\int_{G \cap \Delta_r} v = \sum_{i=0}^\infty v(\lambda_{n_0+i}) (q^{-i}-q^{-i-1}).$$ This is $+\infty$, since it agrees up to a finite number of terms with $$(1-q^{-1})q^{n_0} \sum_{n=1}^\infty v(\lambda_n) q^{-n}.$$ (Substitute $n=n_0+i$.) Hence condition~(2) of the definition of ${\cal G}$ holds. Also, the finite linear combinations of the $\lambda_n$ for $n<n_0$ are coset representatives for $G \cap \Delta_r$ in $G$, so condition~(3) of the definition of ${\cal G}$ follows from properties (2) and~(3) of a decreasing basis. Hence $G \in {\cal G}$. \medskip Conversely, suppose $G \in {\cal G}$. Since $G \cap \Delta_r$ is compact, only zero can be a limit point of $\{|x|:x \in G \cap \Delta_r\}$. Combined with condition~(3) of the definition of ${\cal G}$, this implies that the nonzero absolute values of elements of $G$ consist of a decreasing sequence $\{r_n\}_{n \in {\Bbb Z}}$, possibly terminating in either direction. For each $r_n$, choose an ${{\Bbb F}_q}$-basis for $(G \cap \Delta_{r_n})/(G \cap \Delta_{r_n}')$, which is finite, by condition~(1) in the definition of ${\cal G}$ when $r_n \le r$, and by condition~(3) when $r_n>r$. We claim that the concatenation of these bases (lifted to $C$) is a descending basis for $G$. Properties (1) and~(2) of a descending basis are clearly satisfied. Property~(3) follows from condition~(3) in the definition of ${\cal G}$, and property~(4) is equivalent to condition~(2) in the definition of ${\cal G}$, as in the proof of the converse above. \medskip \noindent {\em Proof of (1), (3), (4), and (5): } Part~(1) follows trivially from~(2). Part~(3) follows since if $G \in {\cal G}$, the open subgroup $G \cap \Delta_r$ is compact. For part~(4), we may assume by~(1) that $r$ is smaller than the absolute value of the smallest nonzero element of $G$, and see what the conditions in the definition of ${\cal G}$ say in this case. Similarly, part~(5) follows by assuming $r$ is greater than the absolute value of all elements of $G$. \medskip \noindent {\em Proof of (6): } In constructing a descending basis for $G$ as in the proof of~(2), we may assume that the chosen basis for $(G \cap \Delta_{r_n})/(G \cap \Delta_{r_n}')$ contains a basis for its subspace $(H \cap \Delta_{r_n})/(H \cap \Delta_{r_n}')$. It is then easy to check that the concatenation of the latter bases (lifted to $C$) is a descending basis for $H$, so that $H \in {\cal G}$ by~(2). \end{pf} \section{Maps and Fibers} \label{kernels} Here we state basic results on the maps on $C$ defined by elements of ${\cal P}$ or ${\cal F}$, and in particular study their fibers. The proofs will be postponed until the next section. First we have a fractional version of ``Picard's Theorem,'' which in the non-archimedean situation asserts that non-constant everywhere convergent power series are surjective. The openness of the map is a kind of local surjectivity. \begin{theorem} \label{nonlinearsurjective} If $f \in {\cal P}$ is not a constant series, then $f:C \rightarrow C$ is a surjective open map. \end{theorem} This theorem has two easy consequences. \begin{cor} \label{ffdomain} If $f,g \in {\cal P}$ are not constant, then $f \circ g$ is not constant. In particular, the noncommutative ring ${\cal F}$ has no zero divisors. \end{cor} \begin{cor} \label{distinct} Distinct series of ${\cal P}$ yield distinct maps on $C$. \end{cor} \begin{theorem} \label{locallycompact} If $f = \sum_{e \in E} a_e z^e \in {\cal P}$ is not constant, then for each $c \in C$, $f^{-1}(c)$ is a locally compact subset of $C$, and is compact if and only if there is a largest $e$ for which $a_e \not=0$. If $f = \sum_{n \in {\Bbb Z}} a_n \tau^n \in {\cal F}$ is nonzero, then $G \stackrel{\text{def}}{=} \ker f$ is in ${\cal G}$. Moreover, \begin{center} $G$ is discrete $\iff$ $a_n=0$ for $n \ll 0$. \\ $G$ is compact $\iff$ $a_n=0$ for $n \gg 0$. \end{center} \end{theorem} \noindent (The last two statements are dual in the sense of Section~\ref{adjointpd}.) We have a converse for $f \in {\cal F}$: \begin{theorem} \label{existence} Each $G \in {\cal G}$ is the kernel of some $f \in {\cal F}$. The $f$ is essentially unique: any $g \in {\cal F}$ with the same kernel is of the form $\epsilon \tau^n \circ f$ for some $\epsilon \in C^\ast$ and $n \in {\Bbb Z}$. (In other words, by Corollary~\ref{units} below, $g$ differs from $f$ only by a unit of ${\cal F}$.) \end{theorem} Because $C$ is not locally compact, Theorem~\ref{locallycompact} implies it is not the union of $f^{-1}(0)$ and $g^{-1}(0)$ for nonzero $f,g \in {\cal P}$. This proves the following. \begin{cor} \label{ppdomain} The commutative ring ${\cal P}$ has no zero divisors. \end{cor} Recall that for a power series $f(z) = \sum_{i=0}^\infty a_i z^i \in C[[z]]$, the Newton polygon is the lower convex hull of the set of points $(i,v(a_i))$ in the plane, and that the Newton polygon gives information on the valuations of the zeros of $f$. (See~\cite{amice},~\cite{artin},~\cite{kneser},~\cite{koblitz}, or~\cite{neukirch}.) Similarly, we define the Newton polygon of $f=\sum_{e \in E} a_e z^e \in {\cal P}$ to be the lower convex hull of the set of points $(e,v(a_e))$ in the plane. Let $v_p$ denote the $p$-adic valuation on $\Bbb Q$. For $f = \sum_{e \in E} a_e z^e \in {\cal P}$, and $n \ge 0$, define \begin{equation} \label{fndefinition} f_n = \sum_{v_p(e) \ge -n} a_e z^e. \end{equation} Therefore $f_n$ is the $p^n$-th root of a power series. Since $f$ is convergent everywhere, we see that the $f_n$ converge to $f$ uniformly on bounded subsets (hence also in ${\cal P}$) as $n$ tends to infinity. \begin{theorem} \label{nonlinearnewton} Suppose $f \in {\cal P}$ is not constant, and $c \in C$. Then there is a canonical measure $\mu$ on $f^{-1}(c)$, characterized by $$\mu \left( f^{-1}(c) \cap (\alpha+\Delta_r) \right) = \lim_{n \rightarrow \infty} \frac{\text{\# zeros of $(f_n-c)^{p^n}$ in $(\alpha+\Delta_r)$}}{p^n},$$ for all $\alpha \in C$, $r>0$. (Here zeros are to be counted with multiplicity.) The horizontal length of the segment of the Newton polygon of $f-c$ having slope $s$ (if any) equals $$\mu(\{\,z \in f^{-1}(c) \mid v(z)=-s \,\}).$$ Also, $$\mu(f^{-1}(c)) = \sup \{\,e \mid a_e \not=0 \,\}.$$ (Both might be $+\infty$.) If $f \in C[[z]]$, then $\mu$ is the ``counting measure''; i.e., $\mu(\{x\})$ is the multiplicity of $x$ as a zero of $f-c$. Finally, if $f \in {\cal F}$ is nonzero and $c=0$, then $\mu$ is a Haar measure on the locally compact group $\ker f$. \end{theorem} \begin{cor} \label{trivialkernel} If $f \in {\cal P}$ vanishes only at 0, then $f(z)=a z^e$ where $a \in C^\ast$ and $e \in E$. \end{cor} \begin{pf} By Theorem~\ref{nonlinearnewton}, any series with at least two monomials has a ero of nonzero valuation. \end{pf} \begin{cor} \label{units} The unit group of ${\cal F}$ consists of the elements of the form $a \tau^n$ where $a \in C^\ast$ and $n \in {\Bbb Z}$. \end{cor} \begin{pf} Any unit of ${\cal F}$ must have trivial kernel, and so must be of the form in Corollary~\ref{trivialkernel}, with the exponent a power of $q$. On the other hand, it is easy to write down inverses of elements of this form. \end{pf} \begin{theorem} \label{factors} Let $f,g$ be nonzero elements of ${\cal F}$. Then $\ker g \subseteq \ker f$ if and only if there exists $h \in {\cal F}$ such that $f=h \circ g$. In this case, the $h$ is unique, and $\ker h = g(\ker f)$. \end{theorem} \section{Proofs} \label{proofs} Here we provide proofs of the theorems stated in the previous section. \begin{lemma} \label{mean} Suppose $f(z)=\sum_{v_p(e) \ge -n} a_e z^e \in {\cal P}$, and $j>0$ is such that $a_j \not= 0$. Then there exists $x \in C$ such that $f(x)=0$ and $|x| \le |f(0)/a_j|^{1/j}$. \end{lemma} \begin{pf} This follows from looking at the Newton polygon for the (ordinary) power series $f(z)^{p^n}.$ \end{pf} Let $f_n$ be as in~(\ref{fndefinition}). \begin{lemma} \label{cosets} Suppose $f \in {\cal P}$ is not a constant series. Fix $r,R \in {\Bbb R}$, with $0<r<R$, and fix $c \in C$. Then the images of $f^{-1}(c) \cap \Delta_R$ and $f_n^{-1}(c) \cap \Delta_R$ in $\Delta_R/\Delta_r$ coincide for $n \gg 0$. \end{lemma} \begin{pf} By considering $f-c$, we may assume $c=0$. By composing with scalar multiplcations on both sides, we may assume $R=1$, and that the largest coefficient of $f$ other than the constant coefficient is of absolute value 1. Let $a_e z^e$ be the term of $f$ with largest $e$ for which $|a_e|=1$. Let $\epsilon_n=\sup|(f_{n+1}-f_n)(\Delta_R)|$. Since $f_n \rightarrow f$, $\epsilon_n \rightarrow 0$. Suppose $\lambda_n \in f_n^{-1}(0) \cap \Delta_R$. Then $$|f_{n+1}(\lambda_n)| = |(f_{n+1}-f_n)(\lambda_n)| \le \epsilon_n.$$ Provided that $n+1 \ge -v_p(e)$, the coefficient of $z^e$ in $f_{n+1}(z+\lambda_n)$ has absolute value 1, by choice of $e$ and since $|\lambda_n| \le 1$. Hence by Lemma~\ref{mean}, there exists $z_n$ such that $|z_n| \le \epsilon_n^{1/e}$ and $$f_{n+1}(z_n+\lambda_n)=0.$$ Let $\lambda_{n+1}=z_n+\lambda_n$. Provided $n$ was chosen large enough, we can repeat the argument to construct a sequence $\lambda_n,\lambda_{n+1},\lambda_{n+2},\ldots$ with $\lambda_i \in f_i^{-1}(0)$ and $$|\lambda_{i+1}-\lambda_i| \le \epsilon_i^{1/e} \le r.$$ Since $\epsilon_i \rightarrow 0$, the $\lambda_i$ converge to a limit $\lambda_\infty$ with $|\lambda_\infty-\lambda_n| \le r$. Also since $f_n \rightarrow f$ uniformly on $\Delta_R$, $$f(\lambda_\infty)=\lim_{i \rightarrow \infty} f_i(\lambda_i)=0.$$ For the other direction, suppose $\lambda \in f^{-1}(0) \cap \Delta_R$. By Lemma~\ref{mean}, for $n \gg 0$, there exists $z_n$ such that $f_n(z_n+\lambda)=0$ and $$|z_n| \le |f_n(\lambda)|^{1/e} \rightarrow 0$$ as $n \rightarrow \infty$. In particular, for sufficiently large $n$ (and how large does not depend on $\lambda$), $|z_n| \le r$, so $z_n+\lambda \in f_n^{-1}(0)$ is in the same coset of $\Delta_r$ as $\lambda$. \end{pf} From the preceding we can deduce two useful corollaries. First of all, we see that Lemma~\ref{mean} extends to arbitrary non-constant $f \in {\cal P}$. \begin{cor} \label{supermean} Suppose $f(z)=\sum_{e \in E} a_e z^e \in {\cal P}$, and $j>0$ is such that $a_j \not= 0$. Then there exists $x \in C$ such that $f(x)=0$ and $|x| \le |f(0)/a_j|^{1/j}$. \end{cor} Next we have a finiteness result, which will be used to prove that fibers are locally compact. \begin{cor} \label{preimage} If $f \in {\cal P}$ is non-constant, $0<r<R$, and $c \in C$, then the image of $f^{-1}(c) \cap \Delta_R$ in $\Delta_R/\Delta_r$ is finite. \end{cor} \begin{pf} The result with $f$ replaced by $f_n$ is proved by applying the theory of Newton polygons to the power series $f_n^{p^n}-c^{p^n}$. Now apply Lemma~\ref{cosets}. \end{pf} \begin{pf*}{Proof of Theorem~\ref{nonlinearsurjective}} Applying Corollary~\ref{supermean} to $f-c$ where $f \in {\cal P}$ is not a constant series and $c \in C$ shows that there is a solution to $f(z)-c=0$, so $f$ is surjective. In checking that $f$ defines an open map, we may reduce to proving $f$ is open at 0, by composing $f$ with a translation. Moreover we may assume $f(0)=0$. If $c \in C$ is small, Corollary~\ref{supermean} applied to $f-c$ shows that $f(z)-c=0$ has a solution near 0. This is exactly what is needed to prove that $f$ is open at 0. \end{pf*} \begin{pf*}{Proof of Theorems \ref{locallycompact} and~\ref{nonlinearnewton}} First let us show that for any $r>0$, $f^{-1}(c) \cap \Delta_r$ is compact. Suppose $\alpha_1,\alpha_2,\ldots$ is a sequence in $f^{-1}(c) \cap \Delta_r$. By Corollary~\ref{preimage}, these elements lie in finitely many cosets of $\Delta_{r/2}$, so we can find an infinite subsequence within one coset. Next we can find a subsequence of this subsequence lying within a single coset of $\Delta_{r/3}$, and so on, with the $n$-th subsequence lying within a coset of $\Delta_{r/n}$. By diagonalization we obtain a convergent subsequence of the original sequence. Since $f$ is continuous, $f^{-1}(c)$ is closed (and so is $\Delta_r$), so the limit lies in $f^{-1}(c) \cap \Delta_r$. Thus $f^{-1}(c) \cap \Delta_r$ is compact. Taking larger and larger $r$ shows that $f^{-1}(c)$ is locally compact. \medskip Before completing the proof of Theorem~\ref{locallycompact}, let us turn to Theorem~\ref{nonlinearnewton}. First we check that the limit in the definition of $\mu$ exists. Without loss of generality we may assume $c=0$. Also we may assume $\alpha=0$, by considering $f(z+\alpha)$. Let $\ell_r(f)$ denote the total horizontal length of the segments of the Newton polygon of $f$ whose slope is less than $\log r$, which is the largest $x$-coordinate of a point of contact of the Newton polygon with a supporting line of slope $\log r$. The fact that $f$ is everywhere convergent implies that below any line in ${\Bbb R}^2$ there are at most finitely vertices, so for each $r$, $\ell_r(f_n)=\ell_r(f)$ for $n \gg 0$. The Newton polygon for $f_n^{p^n}$ is the dilation of that of $f_n$ by the factor $p^n$, so $\ell_r(f_n^{p^n})=p^n \ell_r(f_n)$. The theory of Newton polygons for ordinary power series implies that $\ell_r(f_n^{p^n})$ counts the number of zeros of $f_n^{p^n}$ (with multiplicity) in $\Delta_r$, so the limit in the definition of $\mu$ (for $c=\alpha=0$) converges to $\ell_r(f)$. To check that $\mu$ is truly a measure, it will suffice to check that the definition is consistent in the sense that when $r<R$, its value on $f^{-1}(c) \cap (\beta+\Delta_R)$ equals the sum of its values on $f^{-1}(c) \cap (\alpha+\Delta_r)$ with $\alpha+\Delta_r$ ranging over the cosets of $\Delta_r$ contained in $\beta+\Delta_R$. By Lemma~\ref{cosets} and Corollary~\ref{preimage}, we need concern ourselves with only finitely many of these cosets. Now the result is clear from the finite additivity of the right hand side of the definition of $\mu$. By what we have shown so far, $$\mu(\{\, z \in f^{-1}(0) : |z| \le r \,\}) = \ell_r(f).$$ Taking this and subtracting the same with $r$ replaced by $r-\epsilon$ (with $s=\log r$, and $\epsilon$ sufficiently small) shows that $$\mu(\{\,z \in f^{-1}(c) \mid v(z)=-s \,\})$$ equals the horizontal length of the Newton polygon segment of slope $s$. We now check the final statements of Theorem~\ref{nonlinearnewton}. By what we have shown so far, $\mu(f^{-1}(c))$ is the sum of the horizontal lengths of all the segments, and this is clearly $\sup \{\,e \mid a_e \not=0 \,\}$. If $f \in C[[z]]$, the theory of Newton polygons for power series tells us that the lengths of the horizontal segments are counting zeros, so $\mu$ is the counting measure. Finally, if $f \in {\cal F}$ is nonzero and $c=0$, it is clear from the definition that $\mu$ is translation-invariant on the locally compact group $\ker f$. Moreover, $\mu(\ker f)>0$ by the formula we just derived for the measure of the whole space. Thus $\mu$ is a Haar measure on $\ker f$. This completes the proof of Theorem~\ref{nonlinearnewton}. \medskip We now resume the proof of Theorem~\ref{locallycompact}. First, if there is a largest $e$ for which $a_e \not=0$, then the slopes of the Newton polygon of $f-c$ are bounded above for each $c \in C$, and hence by Theorem~\ref{nonlinearnewton}, $f^{-1}(c) \subset \Delta_r$ for some $r>0$. Thus $f^{-1}(c)=f^{-1}(c) \cap \Delta_r$, which we showed already was compact. On the other hand, if there is no largest $e$ for which $a_e \not=0$, then the Newton polygon of $f-c$ has infinitely many segments of increasing slope, so there is a sequence of elements in $f^{-1}(c)$ of decreasing valuation. Such a sequence cannot have a convergent subsequence, so $f^{-1}(c)$ is not compact. From now on, we assume $f=\sum_{n \in {\Bbb Z}} a_n \tau^n \in {\cal F}$, $c=0$, and $G=\ker f$. By the first paragraph of this proof, $G \cap \Delta_r$ is compact for any $r>0$, proving that $G$ satisfies condition~(1) in the definition of ${\cal G}$. Also by Corollary~\ref{preimage}, the image of $\ker f \cap \Delta_{nr}$ in $\Delta_{nr}/\Delta_r$ is finite for all $n$, so condition (3) in the definition of ${\cal G}$ follows. Now we check condition~(2); i.e., that if $r>0$, then $\int_{G \cap \Delta_r} v =+\infty$. If we integrate only over the subset $$\{\, z \in G \mid v(z)=-s \,\}$$ we get $-s$ times the Haar measure of this set, which, by what we just proved, is the same as $-s$ times the horizontal length of the segment of the Newton polygon of slope $s$, which is the {\em vertical} displacement as one moves along the segment from right to left. If we sum over all the segments of slopes less than $\log r$, we deduce that $\int_{G \cap \Delta_r} v$ equals the vertical displacement as one moves along the section of the Newton polygon to the left, starting at some point depending on $r$. This vertical displacement is $+\infty$, since $v(a_n) \rightarrow +\infty$ as $n \rightarrow -\infty$. Thus $G \in {\cal G}$. Finally, we check the last equivalences in Theorem~\ref{locallycompact}. First, $G$ is discrete if and only if there are no zeros of large finite positive valuation. By Theorem~\ref{nonlinearnewton}, this happens if and only if $a_n=0$ for $n \ll 0$. Lastly, the criterion for $G$ to be compact follows from our earlier criterion for fibers of general elements of ${\cal P}$. \end{pf*} For future reference, we record the following well-known result. \begin{lemma} \label{known} If the zeros of a separable polynomial $f(z) \in C[z]$ form a sub-${{\Bbb F}_q}$-vector space of $C$, then $f \in C\{\tau\}$. \end{lemma} \begin{pf} Proposition~1.3 in~\cite{DH1} proves this for $q=p$, so $f$ is an additive polynomial. Comparing zeros and linear coefficients shows that $f(cz)=cf(z)$ for all $z \in {{\Bbb F}_q}$, and this forces $f \in C\{\tau\}$. \end{pf} \begin{lemma} \label{image} If $G \in {\cal G}$ and $f \in {\cal F}$ has kernel $G \cap \Delta_r$ for some $r>0$, then $f(G)$ is in ${\cal G}$ and is discrete. \end{lemma} \begin{pf} Since $f$ has compact kernel, its coefficient of $\tau^n$ is zero for $n \gg 0$ by Theorem~\ref{locallycompact}. Hence in the series for $f(z)$, the last term dominates for large $z$. Thus $|f(z)| \rightarrow \infty$ as $|z| \rightarrow \infty$. Let $g_1,g_2,\ldots$ be representatives for the cosets of $G \cap \Delta_r$. These form a sequence tending to infinity, by condition (3) in the definition of ${\cal G}$. Then $$f(G)=\{f(g_1),f(g_2),\ldots\}$$ also consists of a sequence tending to infinity, so it is in ${\cal G}$ and is discrete. \end{pf} \begin{pf*}{Proof of Theorem~\ref{existence}, existence} $\left.\right.$ \noindent{\em Case 1:} $G$ discrete. \noindent Then by Lemma~\ref{known}, we may take $$f(z) = z \prod_{g \in G, g \not=0} (1-z/g).$$ \noindent{\em Case 2:} $G$ compact. \noindent We may assume $G$ is infinite, since otherwise we are in Case 1. Without loss of generality suppose $G \subset \Delta_1$. (If for some small $c \in C$ we can get $cG$ as a kernel of $f$, then we can get $G$ as the kernel of $f(cz)$.) By Proposition~\ref{class}, there exists a descending basis $\lambda_1,\lambda_2,\ldots$ for $G$. (It terminates on the left by condition~(3) of the definition of a descending basis.) Let $V_n$ be the ${{\Bbb F}_q}$-vector space generated by $\lambda_1,\ldots,\lambda_n$. Define $$g_n(z) = \prod_{\lambda \in V_n} (z-\lambda), \qquad \qquad h_n(z) = g_n(z)^{1/q^n}.$$ By Lemma~\ref{known}, $g_n \in C\{\tau\}$, so $h_n \in {\cal F}^-$, and both have coefficients bounded in absolute value by 1, since $G \subset \Delta_1$. We will eventually show that the $h_n$ converge to the desired $f \in {\cal F}^-$. We have $$g_{n+1}(z) = g_n(z)^q - g_n(\lambda_{n+1})^{q-1} g_n(z)$$ since both sides are monic ${{\Bbb F}_q}$-linear polynomials of degree $q^{n+1}$ which vanish on $V_{n+1}$. Raising to the $1/q^{n+1}$ power yields \begin{equation} \label{recurrence} h_{n+1}(z) = h_n(z) - c_n h_n(z)^{1/q} \end{equation} where $c_n=h_n(\lambda_{n+1})^{1-1/q}$. We claim that $|c_n| \rightarrow 0$ as $n \rightarrow \infty$. We have \begin{eqnarray*} v(h_n(\lambda_{n+1})) & = & 1/q^n \sum_{\lambda \in V_n} v(\lambda_{n+1}-\lambda) \\ & \ge & 1/q^n \sum_{\lambda \in V_n} \min\{v(\lambda_{n+1}),v(\lambda)\} \\ & \ge & 1/q^n \sum_{\lambda \in V_n, \lambda \not= 0} v(\lambda), \end{eqnarray*} since $v(\lambda_{n+1}) \ge v(\lambda)$ for all nonzero $\lambda \in V_n$, by the definition of descending basis. As $n \rightarrow \infty$, this tends to $\int_G v = +\infty$. Hence $$|c_n| = \exp(-(1-1/q) v(h_n(\lambda_{n+1}))) \rightarrow 0.$$ By~(\ref{recurrence}), we see that $$\|h_{n+1}(z) - h_n(z)\| = \|c_n h_n(z)^{1/q}\| \le |c_n| \rightarrow 0$$ as $n \rightarrow \infty$. Hence the $h_n$ converge to some $f \in {\cal F}$. If $z \in V_n$, then $h_i(z)=0$ for $i \ge n$, so $f(z)=0$. Thus $\ker f$ contains $\bigcup_{n=1}^\infty V_n$ and its closure, which is $G$. On the other hand, if $z \not\in G$, and we set $$\delta = \inf_{\lambda \in G} |z-\lambda|,$$ then directly from the definitions of $g_n(z)$ and $h_n(z)$ we get $$|g_n(z)| \ge \delta^{q^n}, \qquad \qquad |h_n(z)| \ge \delta,$$ for all $n$, so in particular, $f(z) = \lim_{n \rightarrow \infty} h_n(z) \not= 0$. Hence $\ker f = G$, as desired. \medskip \noindent{\em Case 3:} $G \in {\cal G}$ arbitrary. \noindent By Case 2, we can find $g \in {\cal F}$ with kernel $G \cap \Delta_1$. By Lemma~\ref{image}, $g(G)$ is in ${\cal G}$ and is discrete, so by Case 1, there exists $h \in {\cal F}$ with kernel $g(G)$. Let $f=h \circ g$. Then $$f(x)=0 \iff h(g(x))=0 \iff g(x) \in g(G) \iff x \in G,$$ since $\ker g \subseteq G$. \end{pf*} Before proving Theorem~\ref{factors} and the uniqueness part of Theorem~\ref{existence}, let us prove the following ``remainder'' lemma. \begin{lemma} \label{remainder} If $f \in {\cal F}$, then there exists $q \in {\cal F}$ such that $$f = q \circ (1 - \tau) + f(1).$$ \end{lemma} \begin{pf} Suppose $f = \sum_{n \in {\Bbb Z}} a_n \tau^n \in {\cal F}$. Without loss of generality, assume $f(1)=\sum_{n \in {\Bbb Z}} a_n =0$. Then the rate of growth of the $a_n$ imposed by Proposition~\ref{convergence} implies that if $$b_n \stackrel{\text{def}}{=} \sum_{m=-\infty}^n a_m = -\sum_{m=n+1}^\infty a_m,$$ then $q \stackrel{\text{def}}{=} \sum_{n \in {\Bbb Z}} b_n \tau^n$ belongs to ${\cal F}$. (Use the first definition of $b_n$ to get convergence on the left, and the second for convergence on the right.) Now $$q \circ (1-\tau) = \sum_{n \in {\Bbb Z}} (b_n-b_{n-1}) \tau^n = \sum_{n \in {\Bbb Z}} a_n \tau^n = f.$$ \end{pf} \begin{pf*}{Proofs of Theorem~\ref{factors} and the uniqueness in Theorem~\ref{existence}} In Theorem~\ref{factors}, it is clear that if $f=h \circ g$, then $\ker g \subseteq \ker f$, so we will concern ourselves with converse, that if $\ker g \subseteq \ker f$, then there exists $h \in {\cal F}$ such that $f = h \circ g$. Then the uniqueness of $h$ is clear from Corollary~\ref{ffdomain}, and $\ker h$ must be $g(\ker f)$ by the surjectivity of $g$ from Theorem~\ref{nonlinearsurjective}. \noindent{\em Step 1: } Prove Theorem~\ref{factors} for $g= \alpha_0 \circ (1-\tau) \circ \alpha_1 \circ (1-\tau) \circ \alpha_2 \cdots (1-\tau) \circ \alpha_d$, where $\alpha_0,\alpha_1,\ldots,\alpha_d \in C^\ast$. \medskip \noindent We use induction on $d$. The base case $d=0$ is trivial. Suppose $d \ge 1$. Let $\tilde{g} = \alpha_1 \circ (1-\tau) \circ \alpha_2 \cdots (1-\tau) \circ \alpha_d$. Then the inductive hypothesis implies that $f = j \circ \tilde{g}$ for some $j \in {\cal F}$. Since $f$ kills $\ker g$, $j$ kills $\tilde{g}(\ker g) = \ker( \alpha_1 \circ (1-\tau)) = {{\Bbb F}_q}$. In particular $j$ kills 1, so by Lemma~\ref{remainder}, $j=q\circ (1-\tau)$ for some $q$. If we now let $h=q\circ \alpha_0^{-1}$, then $f=h \circ g$. \medskip \noindent{\em Step 2: } Every separable $g \in C\{\tau\}$ of $\tau$-degree $d$ is of the form in Step 1. \medskip \noindent Again we use induction on $d$. The base case $d=0$ is trivial. If $d \ge 1$, there exists $c \in \ker g$, $c \not=0$. Then $g$ kills $\ker ((1-\tau) \circ c^{-1})$, so $g = h \circ (1-\tau) \circ c^{-1}$ for some $h \in {\cal F}$ by Step 1, and the proof of the existence of $h$ in fact shows that $h$ must be a separable element of $C\{\tau\}$. Applying the inductive hypothesis to $h$ produces the desired factorization. \medskip \noindent{\em Step 3: } Theorem~\ref{factors} holds when $g \in C\{\tau,\tau^{-1}\}$. \medskip \noindent Let $\tau^n$ be the lowest (most negative) power occuring in $g$. Then by Steps 1 and 2, we can write $f=h' \circ (\tau^{-n} \circ g)$, so we can take $h=h' \circ \tau^{-n}$. \medskip \noindent{\em Step 4: } If $g_n,h_n \in {\cal F}$ are nonzero, $f = h_n \circ g_n$ for each $n$, and the $g_n$ converge to some nonzero $g \in {\cal F}$, then the $h_n$ converge to some $h \in {\cal F}$ and $f = h \circ g$. \medskip \noindent Fix $x \in C$. We have $h_n(x)=f(g_n^{-1}(x))$ where $g_n^{-1}(x)$ denotes any $y_n \in C$ such that $g_n(y_n)=x$. (By Theorem~\ref{nonlinearsurjective}, such $y_n$ exist.) We will construct a sequence of such $y_n$ which {\em converges}. Fix a nonzero term $a_j \tau^j$ in $g$. Choose some large $R>0$, then choose some large $n_0>0$. (We'll specify how large as we go along; how large we need $n_0$ to be depends on how large $R$ was taken to be.) If $n \gg n_0$, the coefficient of $\tau^j$ in $g_n$ has the same absolute value as $a_j$, provided $n_0$ was chosen large enough. Then by Corollary~\ref{supermean}, there exists a solution $y_{n_0} \in \Delta_R$ to $g_{n_0}(y_{n_0})=x$, if $R$ was chosen large enough. Now inductively define $\epsilon_n$ and $y_{n+1} \in \Delta_R$ for $n \ge n_0$ as follows. Pick $\epsilon_n$ such that $g_{n+1}(\epsilon_n)=(g_n-g_{n+1})(y_n)$, and set $y_{n+1}=y_n+\epsilon_n$. Since the $g_n$ converge uniformly on $\Delta_R$, and since $y_n \in \Delta_R$ by the inductive hypothesis, $(g_n-g_{n+1})(y_n)$ can be assumed to be arbitrary small, if $n_0$ was chosen large enough. Then by Corollary~\ref{supermean}, $\epsilon_n \in \Delta_R$, $y_{n+1} \in \Delta_R$, and \begin{eqnarray*} g_{n+1}(y_{n+1}) & = & g_{n+1}(y_n) + g_{n+1}(\epsilon_n) \\ & = & g_{n+1}(y_n) + g_n(y_n) - g_{n+1}(y_n) \\ & = & x. \end{eqnarray*} Moreover Corollary~\ref{supermean} guarantees that the $\epsilon_n$ can be chosen going to zero, so the $y_n$ converge as desired. Since $f$ is continuous, we see that the sequence $h_n(x)$ converges. In fact, as is clear from the construction of the $y_n$, the rate of convergence depends not on $x$, but only on an upper bound for $|x|$, so the $h_n$ converge uniformly on every bounded subset of $C$. By Theorem~\ref{topologicalring}, the $h_n$ converge to some $h \in {\cal F}$. Taking the limit of $f=h_n \circ g_n$ as $n \rightarrow \infty$ yields $f=h \circ g$. \medskip \noindent{\em Step 5: } Uniqueness holds in Theorem~\ref{existence} when $G \in {\cal G}$ is compact or discrete. Suppose $G \in {\cal G}$ is compact. By the proof of existence part of Theorem~\ref{existence}, there is a sequence of elements $j_n \in C\{\tau,\tau^{-1}\}$ with $\ker j_n \subseteq G$, which converges to an element $j \in {\cal F}$ with kernel $G$. Let $f$ be any other element of ${\cal F}$ with kernel $G$. By Step 3, we may write $f=h_n \circ j_n$, and then by Step 4, we may write $f=h \circ j$. But $j$ is surjective by Theorem~\ref{nonlinearsurjective}, so $\ker h=j(\ker f)=0$ and hence $h$ is a unit by Corollaries \ref{trivialkernel} and~\ref{units}. If $G \in {\cal G}$ is discrete, we can find a sequence $j_n \in C\{\tau\}$ with $\ker j_n \subseteq G$, which converges to an element $j \in {\cal F}$ with kernel $G$, so the same proof works. \medskip \noindent{\em Step 6: } Theorem~\ref{factors} holds when $\ker g$ is compact or discrete. Indeed, by Step 5, $g$ must be a limit of elements $g_n \stackrel{\text{def}}{=} h \circ j_n \in C\{\tau,\tau^{-1}\}$ with $\ker g_n = \ker j_n \subseteq \ker g$, and we can then apply Step 4. \medskip \noindent{\em Step 7: } Theorem~\ref{factors} holds. By the existence part of Theorem~\ref{existence}, we can find $g_1 \in {\cal F}$ with compact kernel $G \cap \Delta_1$. By Step 6, we can write $g=j \circ g_1$, and $\ker j = g_1(\ker g)$, which is discrete by Lemma~\ref{image}. By Step 6 again, we can write $f = h_1 \circ g_1$, and $$\ker j = g_1(\ker g) \subseteq g_1(\ker f) = \ker h_1.$$ By Step 6 yet a third time, we can write $h_1 = h \circ j$. Then $$f = h_1 \circ g_1 = h \circ j \circ g_1 = h \circ g.$$ \medskip \noindent{\em Step 8: } Uniqueness holds in Theorem~\ref{existence}. Indeed, suppose $f$ and $g$ are nonzero elements of ${\cal F}$ with the same kernel. By Step 7, each of $f$ and $g$ is a left multiple of the other, so they differ by a unit. \end{pf*} \section{Pontryagin duality for $A$-modules} \label{pda} Here we recall and develop some results on locally compact topological modules to be used later in the paper. Throughout, topological groups, rings, and modules are assumed Hausdorff, and each $\operatorname{Hom}$ consists of continuous homomorphisms and is given the compact-open topology. If $G$ is a locally compact abelian topological group (LCA group), its {\em Pontryagin dual} is $\hat{G} \stackrel{\text{def}}{=} \operatorname{Hom}_{\Bbb Z}(M,{\Bbb R}/{\Bbb Z})$, which is again an LCA group. The elements of $\hat{G}$ are called {\em characters} of $G$. The main theorem of Pontryagin duality is that the natural map $G \rightarrow \Hat{\Hat{G}}$ is an isomorphism. For an account of Pontryagin duality for groups, see~\cite{armacost}. The following technical result on LCA groups is due to Kaplansky (unpublished) and first appears in a paper of Glicksberg~\cite{glicksberg}. See Chapter~10 of~\cite{armacost} for a proof and discussion. \begin{prop} \label{topology} Let $G$ be a group which becomes a LCA group under a topology ${\frak t}$. Let ${\frak t}'$ be a strictly stronger locally compact group topology on $G$. Then there is a ${\frak t}'$-continuous character $\gamma$ of $G$ which is not ${\frak t}$-continuous. \end{prop} \begin{cor} If $f:G \rightarrow H$ is a surjective continuous homomorphism of LCA groups whose dual $\hat{f}:\hat{H} \rightarrow \hat{G}$ is also surjective, then $f$ is a topological isomorphism. \end{cor} \begin{pf} Since $\hat{f}$ is surjective, $f$ is injective by~P.23(b) in~\cite{armacost}. Thus $f$ is an isomorphism, except that $G$ may have a stronger topology than $H$. If it were strictly stronger, then by the proposition above, there would be a character of $G$ not coming from a character of $H$, contradicting the surjectivity of $\hat{f}$. Therefore $f$ is a topological isomorphism. \end{pf} \begin{cor} \label{showdual} Suppose $G$ and $H$ are LCA groups, and there is a continuous bilinear pairing $$G \times H \rightarrow {\Bbb R}/{\Bbb Z}$$ such that the induced maps $$f_1:G \rightarrow \hat{H} \;\;\;\;,\;\;\;\; f_2:H \rightarrow \hat{G}$$ are surjective. Then $f_1$ and $f_2$ are topological isomorphisms. \end{cor} \begin{pf} By a well-known property of the compact-open topology, the induced maps are continuous. By Pontryagin duality (and chasing definitions), $f_2$ is the dual of $f_1$, so $f_1$ is a topological isomorphism by the previous corollary. The symmetric argument works for $f_2$. \end{pf} We will make use of the theory of Pontryagin duality for topological modules developed by Flood~\cite{flood}. Let $A$ be a locally compact commutative topological ring. Consider the class $\cal A$ of locally compact topological $A$-modules. If $M \in \cal A$, then the Pontryagin dual $\hat{M} \stackrel{\text{def}}{=} \operatorname{Hom}_{\Bbb Z}(M,{\Bbb R}/{\Bbb Z})$ of $M$ as a topological group has a natural $A$-module structure, so $\hat{M} \in \cal A$. Let us restate the main theorem of~\cite{flood} for this situation: \begin{theorem}[Pontraygin duality for $A$-modules] \label{pdfora} For each $M \in \cal A$, the topological $A$-modules $\hat{M}$ and $\operatorname{Hom}_A(M,\hat{A})$ are canonically isomorphic, and the canonical map \begin{eqnarray*} M & \rightarrow & \operatorname{Hom}_A(\operatorname{Hom}_A(M,\hat{A}),\hat{A}) \\ m & \mapsto & (f \mapsto f(m)) \end{eqnarray*} is an isomorphism of topological $A$-modules. \end{theorem} If $M,N \in \cal A$, we say $N$ is the Pontryagin dual of $M$ as an $A$-module, if it is isomorphic to $\operatorname{Hom}_A(M,\hat{A})$ as a topological $A$-module. The following corollary is an $A$-module version of Corollary~\ref{showdual}. The next corollary is useful in verifying that two elements of $\cal A$ are Pontryagin duals (as $A$-modules). \begin{cor} \label{showdual2} Suppose $M,N \in \cal A$ and there is a continuous $A$-module pairing $M \times N \rightarrow \hat{A}$ such that the induced maps $$f_1:M \rightarrow \operatorname{Hom}_A(N,\hat{A}), \;\;\;\;\;\;\;\; f_2:N \rightarrow \operatorname{Hom}_A(M,\hat{A})$$ are surjective. Then $f_1$ and $f_2$ are topological $A$-module isomorphisms. \end{cor} \begin{pf} If we compose the pairing with the evaluation-at-1 map $\hat{A} \rightarrow {\Bbb R}/{\Bbb Z}$, we get a pairing $$M \times N \rightarrow {\Bbb R}/{\Bbb Z}$$ and the induced map $M \rightarrow \hat{N}$ is the same as $f_1$ once we identify $\hat{N}$ with $\operatorname{Hom}_A(N,\hat{A})$ as in Theorem~\ref{pdfora}. Similarly $f_2$ is the same as the induced map $N \rightarrow \hat{M}$. In particular, these induced maps are surjective, so by Corollary~\ref{showdual}, they are topological isomorphisms. But we know that $f_1$ and $f_2$ are $A$-module homomorphism as well, so the result follows. \end{pf} \begin{lemma} \label{showdual3} Suppose $M,N$ are locally compact topological ${{\Bbb F}_q}$-vector spaces. If there is a continuous ${{\Bbb F}_q}$-linear pairing $$\langle \;\; , \;\; \rangle : M \times N \rightarrow {{\Bbb F}_q}$$ such that the induced maps $$M \rightarrow \operatorname{Hom}_{{\Bbb F}_q}(N,{{\Bbb F}_q}) \;\;\;\;,\;\;\;\; N \rightarrow \operatorname{Hom}_{{\Bbb F}_q}(M,{{\Bbb F}_q})$$ are surjective, then $N$ is the Pontryagin dual of $M$ as an ${{\Bbb F}_q}$-vector space. If furthermore $A$ is a locally compact ${{\Bbb F}_q}$-algebra and $M$ and $N$ are topological $A$-modules such that $\langle am,n \rangle = \langle m,an \rangle$ for all $a \in A$, $m \in M$ and $n \in N$, then $\operatorname{Hom}_{{\Bbb F}_q}(A,{{\Bbb F}_q}) \cong \hat{A}$ as topological $A$-modules and the map \begin{eqnarray*} [ \;\; , \;\; ] : M \times N & \rightarrow & \operatorname{Hom}_{{\Bbb F}_q}(A,{{\Bbb F}_q}) \cong \hat{A} \\ m \; , \; n \;\; & \mapsto & (a \mapsto \langle am,n \rangle) \end{eqnarray*} is a continuous $A$-module pairing which exhibits $N$ as the Pontryagin dual of $M$ as an $A$-module. \end{lemma} \begin{pf} By counting elements, ${\hat{\Bbb F}_q}$ is a one-dimensional ${{\Bbb F}_q}$-vector space, and we can explicitly identify ${{\Bbb F}_q}$ with ${\hat{\Bbb F}_q}$ by mapping $1$ to the character $\chi=(1/p) \cdot \operatorname{Tr}_{{{\Bbb F}_q}/{{\Bbb F}_p}}$ of ${{\Bbb F}_q}$. Thus the first half of the proposition is just the special case of Corollary~\ref{showdual2} for the discrete ring ${{\Bbb F}_q}$. There is an isomorphism $\operatorname{Hom}_{{\Bbb F}_q}(A,{{\Bbb F}_q}) \cong \hat{A}$ of topological ${{\Bbb F}_q}$-vector spaces by Theorem~\ref{pdfora} (with $M=A$, $A={{\Bbb F}_q}$), and it is easily checked that it preserves the $A$-module structure. As in Corollary~\ref{showdual2}, the pairing $\langle \;\;,\;\; \rangle$ induces a pairing $M \times N \rightarrow {\Bbb R}/{\Bbb Z}$ which identifies $N$ with $\hat{M}$. The condition $\langle am,n \rangle = \langle m,an \rangle$ ensures that this isomorphism is an isomorphism of $A$-modules. Finally, if we identify $N \cong \hat{M}$ with $\operatorname{Hom}_A(M,\hat{A})$ using Theorem~\ref{pdfora}, we get a pairing $$M \times N \rightarrow \hat{A}$$ exhibiting $N$ as the Pontryagin dual of $M$ as an $A$-module, and definition chasing shows that this pairing is $[\;\;,\;\;]$. \end{pf} Now let $X$ be a nonsingular projective curve over ${{\Bbb F}_q}$, let $\infty$ be a closed point, and let $Y=X \setminus \infty$. Let $A$ be the Dedekind ring of regular functions on the affine curve $Y$, and give $A$ the discrete topology. Let $K$ be its fraction field, and let $K_\infty$ be the completion of $K$ at $\infty$. The K\"{a}hler differentials of $A$ over ${{\Bbb F}_q}$ (see~\cite{matsumura}), i.e., the differentials on $X$ which are regular away from $\infty$, form a rank one projective $A$-module $\Omega_A$ inside the one-dimensional $K$-vector space $\Omega_K$ of K\"{a}hler differentials of $K$ over ${{\Bbb F}_q}$. Therefore there is an ideal $J$ of $A$ isomorphic to $\Omega$ as an $A$-module. Let $\Omega_\infty$ be the completion of $\Omega_K$ at $\infty$. This is a one-dimensional vector space over $K_\infty$. By Theorem~3 of Chapter~II in~\cite{weilnt}, any nontrivial character of the locally compact field $K_\infty$ can be used to identify $K_\infty$ with its dual, and it follows that the map \begin{eqnarray} \label{residuepairing} K_\infty \times \Omega_\infty & \rightarrow & {{\Bbb F}_q} \\ a \;\; , \;\; \omega \;\; & \rightarrow & \operatorname{Res}_\infty(a \omega) \nonumber \end{eqnarray} exhibits $K_\infty$ and $\Omega_\infty$ as Pontraygin duals of each other, as topological ${{\Bbb F}_q}$-vector spaces. \begin{theorem} \label{fractionaldual} Let $I$ be a fractional ideal of $A$ (with the discrete topology). Then the pairing~(\ref{residuepairing}) puts $I$ and $\Omega_\infty/(I^{-1} \Omega_A)$ in Pontryagin duality as $A$-modules, so the Pontryagin dual of $I$ as an $A$-module is $K_\infty/(I^{-1}J)$. \end{theorem} \begin{pf} For $\omega \in \Omega_\infty$, we claim that $\omega \in \Omega_A$ if and only if $\operatorname{Res}_\infty(a \omega)=0$ for all $a \in A$. For $n \ge 0$, let \begin{eqnarray*} L(n \infty) & = & \{\, a \in K \mid \operatorname{div}(a) \ge -n \infty \,\} \\ \Omega_{\ge n \infty} & = & \{\, \eta \in \Omega_\infty \mid \operatorname{div}(\eta) \ge n \infty \,\}. \end{eqnarray*} (By the divisor of an element of $\Omega_\infty$ we mean only the part at $\infty$, which is the only part that makes sense.) By the Theorem at the bottom of page~160 in~\cite{eichler}, with $h=0$, ${\frak a}=n \infty$, and $dw$ the principal part system which equals an element of $\Omega_K$ differing from $\omega$ by an element of $\Omega_{\ge n \infty}$ at $\infty$, and equalling zero at other places, we have that $\operatorname{Res}_\infty(a \omega)=0$ for all $a \in L(n \infty)$ if and only if $\omega \in \Omega_A + \Omega_{\ge n \infty}$. But $A = \bigcup_{n \ge 0} L(n \infty)$, so $\operatorname{Res}_\infty(a \omega)=0$ for all $a \in A$ if and only if $$\omega \in \bigcap_{n \ge 0} (\Omega_A + \Omega_{\ge n \infty}) = \Omega_A,$$ since $\Omega_A$ is a discrete subgroup of $\Omega_\infty$, and the $\Omega_{\ge n \infty}$ form a decreasing neighborhood base of 0. Hence $\operatorname{Res}_\infty(a \omega)=0$ for all $a \in I$ if and only if $\omega \in I^{-1} \Omega_A$. By P.22(d) in~\cite{armacost}, the fact that the pairing~(\ref{residuepairing}) puts $K_\infty$ and $\Omega_\infty$ in Pontryagin duality implies that $$\hat{I} \cong \operatorname{Hom}_{{\Bbb F}_q}(I,{{\Bbb F}_q}) \cong \Omega_\infty/(I^{-1} \Omega_A) \cong K_\infty/(I^{-1}J),$$ and it is trivial to check that all our isomorphisms respect the $A$-module structures. \end{pf} Let $A_{\frak q}$, $K_{\frak q}$ denote the completions of $A$, $K$, respectively, at a nonzero prime ${\frak q}$ of~$A$. \begin{theorem} \label{aqdual} The $A_{\frak q}$-module $\operatorname{Hom}_{{\Bbb F}_q}(K_{\frak q}/A_{\frak q},{{\Bbb F}_q})$ is free of rank one. \end{theorem} \begin{pf} Let $t \in K$ be a uniformizing parameter at ${\frak q}$. Then $A_{\frak q} \cong {{\Bbb F}_\qq}[[t]]$, where ${{\Bbb F}_\qq}$ is the residue field at ${\frak q}$. Now for the residue pairing \begin{eqnarray*} {{\Bbb F}_\qq}((t)) \times {{\Bbb F}_\qq}((t))dt & \rightarrow & {{\Bbb F}_q} \\ a \;\;\; , \;\;\; \omega \;\;\;\;\;\;\;\;\;\; & \mapsto & \operatorname{Res}_{\frak q}(a \omega) \end{eqnarray*} which exhibits the locally compact field $K_{\frak q}$ as its own Pontryagin dual, the set of $\omega$ which give zero when paired with any $a \in {{\Bbb F}_\qq}[[t]]$ is exactly ${{\Bbb F}_\qq}[[t]]dt$, by definition of the residue, so by~P.22(d) in~\cite{armacost}, the Pontryagin dual of $A_{\frak q}={{\Bbb F}_\qq}[[t]]$ is $K_{\frak q}/A_{\frak q}={{\Bbb F}_\qq}((t))/{{\Bbb F}_\qq}[[t]]$. \end{pf} \section{Adjoints and Pontryagin duality} \label{adjointpd} The {\em adjoint map} on ${\cal F}$ is the map $f \mapsto f^\ast$ defined by the following proposition. \begin{prop} There is a norm-preserving multiplication-reversing involution \begin{eqnarray*} {\cal F} & \rightarrow & {\cal F} \\ f=\sum_{n \in {\Bbb Z}} a_n \tau^n & \mapsto & f^\ast=\sum_{n \in {\Bbb Z}} {a_n}^{1/q^n} \tau^{-n}. \end{eqnarray*} \end{prop} \begin{pf} The map is well-defined by Proposition~\ref{convergence}, and it is immediate from the definition that $\|f^\ast\|=\|f\|$. One can check $f^{\ast \ast}=f$ and $(f \circ g)^\ast = g^\ast \circ f^\ast$ directly by comparing coefficients, or perhaps more enlighteningly by noting that $$f^\ast = \sum_{n \in {\Bbb Z}} \tau^{-n} a_n$$ is obtained from $f$ by interchanging $\tau$ and $\tau^{-1}$ and reversing the order of multiplication everywhere, and that this preserves the pair of commutation relations $\tau \circ a = a^q \circ \tau$ and $a \circ \tau^{-1} = \tau^{-1} \circ a^q$. \end{pf} We now construct an ${{\Bbb F}_q}$-linear pairing between $\ker f$ and $\ker f^\ast$, for nonzero $f \in {\cal F}$. Suppose $\alpha \in \ker f$ and $\beta \in \ker f^\ast$. Then $f \circ \alpha$ vanishes on 1 (recall that $f \circ \alpha$ denotes the composition of $f$ with the map $x \mapsto \alpha x$), so by Lemma~\ref{remainder}, \begin{equation} \label{def} f \circ \alpha = g_\alpha \circ (1 - \tau) \end{equation} for some $g_\alpha \in {\cal F}$, which is uniquely defined, by Corollary~\ref{ffdomain}. Define $$\langle \alpha,\beta \rangle_f = g_\alpha^\ast(\beta).$$ \begin{prop} If $f \in {\cal F}$ is nonzero, then $$\langle \;\;,\;\; \rangle_f : \ker f \times \ker f^\ast \rightarrow {{\Bbb F}_q}$$ is an ${{\Bbb F}_q}$-vector space pairing. \end{prop} \begin{pf} First let us check that the image lands in ${{\Bbb F}_q}$. Take adjoints of (\ref{def}): \begin{equation} \label{adjdef} \alpha \circ f^\ast = (1 - \tau^{-1}) \circ g_\alpha^\ast. \end{equation} Apply both sides to $\beta$: $$0 = (1 - \tau^{-1}) g_\alpha^\ast(\beta).$$ So $$g_\alpha^\ast(\beta) \in \ker(1 - \tau^{-1}) = {{\Bbb F}_q}.$$ From the definition~(\ref{def}), $g_{\alpha+\beta}=g_\alpha+g_\beta$, so the pairing is linear on the left. Linearity on the right is obvious. \end{pf} We now prove a series of results leading up to Theorem~\ref{duality} below. \begin{lemma} \label{remainder2} If $f \in {\cal F}$, then there exists $q \in {\cal F}$ such that $$f = (1 - \tau) \circ q + f^\ast(1).$$ \end{lemma} \begin{pf} Apply Lemma~\ref{remainder} to $f^\ast$ to get $h$ such that $$f^\ast = h \circ (1-\tau) + f^\ast(1).$$ Then taking adjoints yields \begin{eqnarray*} f & = & (1-\tau^{-1}) \circ h^\ast + f^\ast(1) \\ & = & (1-\tau) \circ (-\tau^{-1}) \circ h^\ast + f^\ast(1), \end{eqnarray*} so we may take $q=-\tau^{-1} \circ h^\ast$. \end{pf} \begin{lemma} \label{oneone} If $f,g \in {\cal F}$ and $f \circ (1-\tau) = (1-\tau) \circ g$, then $f^\ast(1)=g(1)$. \end{lemma} \begin{pf} Use Lemmas \ref{remainder} and~\ref{remainder2} to write \begin{eqnarray*} f & = & (1-\tau) \circ q + f^\ast(1) \\ g & = & h \circ (1-\tau) + g(1). \end{eqnarray*} Substitute these into the given relation: \begin{equation} \label{subst} (1-\tau) \circ q \circ (1-\tau) + f^\ast(1) \circ (1-\tau) = (1-\tau) \circ h \circ (1-\tau) + (1-\tau) \circ g(1). \end{equation} Evaluating the given relation at 1 shows $g(1) \in \ker(1-\tau) = {{\Bbb F}_q}$, so $g(1)$ commutes with $1-\tau$ and we can cancel $1-\tau$ on the right in (\ref{subst}) to obtain $$(1-\tau) \circ q + f^\ast(1) = (1-\tau) \circ h + g(1).$$ Taking adjoints and evaluating at 1 gives the desired result. \end{pf} \begin{prop} \label{backwards} If $f \in {\cal F}$ is nonzero, $\alpha \in \ker f$, and $\beta \in \ker f^\ast$, then $\langle \alpha,\beta \rangle_f = -\langle \beta,\alpha \rangle_{f^\ast}$. \end{prop} \begin{pf} Write \begin{eqnarray*} f \circ \alpha = g_\alpha \circ (1 - \tau) \\ f^\ast \circ \beta = h_\beta \circ (1 - \tau). \end{eqnarray*} Multiply the first by $\beta$ on the left, and take the adjoint of the second and multiply by $\alpha$ on the right: \begin{eqnarray*} \beta \circ f \circ \alpha & = & \beta \circ g_\alpha \circ (1 - \tau) \\ \beta \circ f \circ \alpha & = & (1 - \tau^{-1}) \circ h_\beta^\ast \circ \alpha \\ & = & (1-\tau) \circ (-\tau^{-1}) \circ h_\beta^\ast \circ \alpha. \end{eqnarray*} Equate and apply Lemma~\ref{oneone} to get \begin{eqnarray*} (\beta \circ g_\alpha)^\ast(1) & = & ((-\tau^{-1}) \circ h_\beta^\ast \circ \alpha)(1) \\ g_\alpha^\ast(\beta) & = & - \tau^{-1} (h_\beta^\ast(\alpha)) \\ \langle \alpha, \beta \rangle_f & = & - \tau^{-1}(\langle \beta,\alpha \rangle_{f^\ast}) \\ & = & - \langle \beta,\alpha \rangle_{f^\ast}, \end{eqnarray*} since $\langle \beta,\alpha \rangle_{f^\ast} \in {{\Bbb F}_q}$. \end{pf} \begin{theorem} \label{duality} If $f \in {\cal F}$ is nonzero, then the pairing $$\langle \;\;,\;\; \rangle_f : \ker f \times \ker f^\ast \rightarrow {{\Bbb F}_q}$$ exhibits $\ker f$ as the Pontryagin dual of $\ker f^\ast$ as a topological ${{\Bbb F}_q}$-vector space. \end{theorem} \begin{pf} First we check that the pairing is continuous. Because of Proposition~\ref{backwards}, it will suffice to show that given any bounded subset $B$ of $\ker f^\ast$, there exists a neighborhood $U$ of 0 in $\ker f$ such that $\langle \alpha,\beta \rangle_f=0$ for $\alpha \in U$, $\beta \in B$. Given $\alpha \in \ker f$, write $$f \circ \alpha = g_\alpha \circ (1 - \tau)$$ as in the definition of the pairing. As $\alpha$ tends to zero, each coefficient of $f \circ \alpha$ tends to zero, and so does each coefficient of $g_\alpha$, by the construction in the proof of Lemma~\ref{remainder}. Then the same is true for the coefficients of $g_\alpha^\ast$. Combined with the knowledge that $g_\alpha^\ast$ is everywhere convergent for any $\alpha$, this implies that $g_\alpha^\ast(\beta)$ tends to zero uniformly for $\beta \in B$ as $\alpha$ tends to zero, as desired. Thus the pairing is continuous. The map induced by the pairing is \begin{eqnarray*} \Psi: \ker f & \rightarrow & \operatorname{Hom}_{{\Bbb F}_q}(\ker f^\ast,{{\Bbb F}_q}) \\ \alpha & \mapsto & g_\alpha^\ast|_{\ker f^\ast} \end{eqnarray*} This is well-defined and continuous, since the pairing is. We now show $\Psi$ is surjective. If $\phi \in \operatorname{Hom}_{{\Bbb F}_q}(\ker f^\ast,{{\Bbb F}_q})$ is nonzero, then $H=\ker \phi$ is some open and closed subspace in $G=\ker f^\ast$, so $H \in {\cal G}$ by~(6) in Proposition~\ref{class}. Thus by Theorem~\ref{existence}, there exists $h \in {\cal F}$ such that $\ker h = H$. Since $H$ has codimension 1 in $G$ (as an ${{\Bbb F}_q}$-vector space), $h$ must map $G$ to a one-dimensional ${{\Bbb F}_q}$-vector space inside $C$. By scaling $h$, we may assume $h(G)={{\Bbb F}_q}$. Now $(1 - \tau) \circ h$ has kernel $G$, so by the uniqueness in Theorem~\ref{existence}, $$\tau^n \alpha \circ f^\ast = (1 - \tau) \circ h$$ for some $\alpha \in C^\ast$ and $n \in {\Bbb Z}$. Take adjoints and multiply on the right by $\tau^n$: \begin{eqnarray*} f \circ \alpha & = & h^\ast \circ (1-\tau^{-1}) \circ \tau^n \\ & = & h^\ast \circ (-\tau^{n-1}) \circ (1-\tau). \end{eqnarray*} Evaluating at 1 shows that $\alpha \in \ker f$. Comparing with~(\ref{def}) shows \begin{eqnarray*} g_\alpha & = & h^\ast \circ (-\tau^{n-1}) \\ g_\alpha^\ast & = & - \tau^{1-n} \circ h \\ g_\alpha^\ast(G) & = & - \tau^{1-n}(h(G)) \\ & = & \tau^{1-n}({{\Bbb F}_q}) \\ & = & {{\Bbb F}_q}. \end{eqnarray*} Thus $\phi$ and $\Psi(\alpha)=g_\alpha^\ast|_{\ker f^\ast}$ are nonzero ${{\Bbb F}_q}$-homomorphisms from $G$ to ${{\Bbb F}_q}$ with the same kernel $H$, so $\phi = c \Psi(\alpha) = \Psi(c \alpha)$ for some $c \in {{\Bbb F}_q}^\ast$. Therefore $\Psi$ is surjective. Applying the above argument to $f^\ast$ and invoking Proposition~\ref{backwards} shows that the other induced map $$\ker f^\ast \rightarrow \operatorname{Hom}_{{\Bbb F}_q}(\ker f,{{\Bbb F}_q})$$ is surjective as well. Now apply Lemma~\ref{showdual3}. \end{pf} \begin{rems} If $f = \sum_{i=m}^n a_i \tau^i$ is a finite series with $a_i$ in {\em any} field $L$ of characteristic $p$, then the construction yields a perfect pairing between the kernels of $f$ and $f^\ast$ acting on $\overline{L}$. Here, in addition, the pairing will be $\operatorname{Gal}(L^{\text{sep}}/L)$-equivariant, since its construction is canonical. The same will holds for infinite series, when $L$ has a non-archimedean valuation and $L^{\text{sep}}$ is replaced by topological Galois closure. \end{rems} \medskip Next we prove a compatibility result. \begin{prop} \label{compatible} Let $f$ and $h$ be nonzero elements of ${\cal F}$. Then for all $\alpha \in \ker(f\circ h)$ and $\beta \in \ker f^\ast \subseteq \ker (f \circ h)^\ast$, $$\langle \alpha,\beta \rangle_{f \circ h} = \langle h(\alpha),\beta \rangle_f.$$ Similarly, for all $\alpha \in \ker f \subseteq \ker(h \circ f), \beta \in \ker(h \circ f)^\ast$, $$\langle \alpha,\beta \rangle_{h \circ f} = \langle \alpha,h^\ast(\beta) \rangle_f.$$ \end{prop} \begin{pf} For the first part, write \begin{eqnarray*} (f \circ h) \circ \alpha & = & G_\alpha \circ (1 - \tau) \\ f \circ h(\alpha) & = & g_{h(\alpha)} \circ (1 - \tau) \end{eqnarray*} Subtract to get $$f \circ (h \circ \alpha - h(\alpha)) = (G_\alpha-g_{h(\alpha)}) \circ (1 - \tau).$$ Since $h \circ \alpha - h(\alpha)$ kills $1$, $$h \circ \alpha - h(\alpha) = q \circ (1 - \tau)$$ for some $q \in {\cal F}$, by Lemma~\ref{remainder}. Substitute and cancel $1-\tau$ on the right: $$f \circ q = G_\alpha - g_{h(\alpha)}.$$ Take adjoints and apply both sides to $\beta \in \ker f^\ast$ to get \begin{eqnarray*} 0 & = & G_\alpha^\ast(\beta) - g_{h(\alpha)}^\ast(\beta) \\ & = & \langle \alpha, \beta \rangle_{f \circ h} - \langle h(\alpha),\beta \rangle_f \end{eqnarray*} as desired. For the second part, we have \begin{eqnarray*} \langle \alpha,\beta \rangle_{h \circ f} & = & - \langle \beta, \alpha \rangle_{f^\ast \circ h^\ast} \;\;\; \text{(by Proposition~\ref{backwards})} \\ & = & - \langle h^\ast(\beta), \alpha \rangle_{f^\ast} \;\;\; \text{(by what we just showed)} \\ & = & \langle \alpha,h^\ast(\beta) \rangle_f, \end{eqnarray*} by Proposition~\ref{backwards} again. \end{pf} \begin{prop} \label{annihilator} Suppose $h,g \in {\cal F}$ and $f=h \circ g$. Then $$\{\, \beta \in \ker f^\ast : \langle \alpha,\beta \rangle_f = 0 \;\; \forall \alpha \in \ker g \,\} = \ker h^\ast.$$ \end{prop} \begin{pf} Suppose $\beta \in \ker f^\ast$. Then, by Proposition~\ref{compatible}, \begin{eqnarray*} \langle \alpha,\beta \rangle_f = 0 \;\; \forall \alpha \in \ker g & \iff & \langle \alpha,h^\ast(\beta) \rangle_g = 0 \;\; \forall \alpha \in \ker g \\ & \iff & h^\ast(\beta)=0 \;\;\; \text{ (by Theorem~\ref{duality})}. \end{eqnarray*} \end{pf} Below are three more properties of the pairing. These were discovered by Elkies~\cite{elkies} (for the case of additive polynomials), although some of our proofs are new. We will not need these in the sequel, but they are of interest in their own right. First we explain how the pairing changes if we change the underlying finite field. \begin{prop} Let $q'=q^m$, for some $m \ge 1$. If $f \in {\cal F}$ is actually ${{\Bbb F}_{q'}}$-linear, then we can define another pairing $\langle \;\;,\;\; \rangle'_f$ between $\ker f$ and $\ker f^\ast$, this time taking values in ${{\Bbb F}_{q'}}$. (Note that $f^\ast$ is the same whether $f$ is considered as ${{\Bbb F}_q}$-linear or ${{\Bbb F}_{q'}}$-linear.) Then $\langle \alpha,\beta \rangle_f = \operatorname{Tr}_{{{\Bbb F}_{q'}}/{{\Bbb F}_q}} \langle \alpha,\beta \rangle'_f$. \end{prop} \begin{pf} Write \begin{eqnarray*} f \circ \alpha & = & G_\alpha \circ (1 - \tau^m) \\ & = & G_\alpha \circ (1+ \tau + \cdots + \tau^{m-1}) \circ (1-\tau). \end{eqnarray*} By definition of the pairing, \begin{eqnarray*} \langle \alpha,\beta \rangle_f & = & (G_\alpha \circ (1+ \tau + \cdots + \tau^{m-1}))^\ast(\beta) \\ & = & (1+\tau + \cdots + \tau^{m-1})^\ast(G_\alpha^\ast(\beta)) \\ & = & (1+\tau^{-1} + \cdots + \tau^{-(m-1)})(\langle \alpha,\beta \rangle'_f) \\ & = & \operatorname{Tr}_{{{\Bbb F}_{q'}}/{{\Bbb F}_q}} \langle \alpha,\beta \rangle'_f, \\ \end{eqnarray*} since $\operatorname{Gal}({{\Bbb F}_{q'}}/{{\Bbb F}_q}) = \{ 1, \tau^{-1},\ldots,\tau^{-(m-1)} \}$. \end{pf} \begin{theorem} \label{skewsymmetric} Suppose $f \in {\cal F}$ is nonzero and $f=f^\ast$. Then \begin{enumerate} \item $\langle \alpha,\alpha \rangle_f = 0$ for all $\alpha \in \ker f$. \item If $h \in {\cal F}$ and $f=h^\ast \circ h$, then $\ker h$ is a maximal isotropic closed sub-${{\Bbb F}_q}$-vector space of $\ker f$ under $\langle \;,\; \rangle_f$. \item Conversely, if $H$ is a maximal isotropic closed sub-${{\Bbb F}_q}$-vector space of $\ker f$ under $\langle \;,\; \rangle_f$, then there exists $h \in {\cal F}$ such that $f=h^\ast \circ h$ and $\ker h = H$. \end{enumerate} \end{theorem} \begin{pf} If $p \not=2$, then part~(1) follows already from Proposition~\ref{backwards}. Hence assume $p=2$. Write $f \circ \alpha = g_\alpha \circ (1-\tau)$ as usual. Multiply by $\alpha$ on the left, and apply Lemma~\ref{remainder2} to write \begin{eqnarray*} \alpha \circ f \circ \alpha & = & [(1-\tau) \circ q + (\alpha \circ g_\alpha)^\ast(1)] \circ (1-\tau) \\ & = & (1-\tau^{-1}) \circ j \circ (1-\tau) + g_\alpha^\ast(\alpha) \circ (1-\tau) \;\;\; \text{(where $j=\tau \circ q$).} \end{eqnarray*} Subtract this from its adjoint and use $f=f^\ast$ and $g_\alpha^\ast(\alpha) \in {{\Bbb F}_q}$ to get $$0 = (1-\tau^{-1}) \circ (j^\ast-j) \circ (1-\tau) + g_\alpha^\ast(\alpha) \circ (\tau-\tau^{-1}).$$ Since $\tau-\tau^{-1}=(1-\tau^{-1}) \circ (1-\tau)$ in characteristic 2, we can cancel $1-\tau^{-1}$ on the left and $1-\tau$ on the right to obtain $$0 = (j^\ast-j) + g_\alpha^\ast(\alpha).$$ Take coefficients of $\tau^0$ to deduce $$\langle \alpha,\alpha \rangle_f = g_\alpha^\ast(\alpha) = 0.$$ This proves part~(1). Now, for part~(2), $h$ is continuous, so $\ker h$ is closed. Proposition~\ref{annihilator} says that $$\{\, \beta \in \ker f^\ast : \langle \alpha,\beta \rangle_f = 0 \;\; \forall \alpha \in \ker h \,\} = \ker h,$$ which proves part~(2). Finally let us prove~(3). Assume $H$ is a maximal isotropic closed sub-${{\Bbb F}_q}$-vector space of $\ker f$ under $\langle \;,\; \rangle$. Then $H \in {\cal G}$ by~(6) in Proposition~\ref{class}. By Theorem~\ref{existence} we can find $g \in {\cal F}$ with kernel $H$, and by Theorem~\ref{factors}, we have $f = j \circ g$ for some $j \in {\cal F}$. By Proposition~\ref{annihilator}, $$\ker j^\ast = \{\, \beta \in \ker f^\ast : \langle \alpha,\beta \rangle_f = 0 \;\; \forall \alpha \in H \,\} =H,$$ since $H$ is maximal isotropic. By the uniqueness in Theorem~\ref{existence}, we have $j^\ast = \epsilon \circ \tau^n \circ g$ where $\epsilon \in C^\ast$ and $n \in {\Bbb Z}$. Then $$f = j \circ g = g^\ast \circ \tau^{-n} \circ \epsilon \circ g.$$ Taking adjoints and using $f=f^\ast$ shows that $$f = g^\ast \circ \epsilon \circ \tau^n \circ g.$$ If we equate and cancel $g^\ast$ on the left and $g$ on the right, we obtain $$\tau^{-n} \circ \epsilon = \epsilon \circ \tau^n,$$ which forces $n=0$. Now simply take $h = \sqrt{\epsilon} \circ g$ where $\sqrt{\epsilon}$ is any square root of $\epsilon$ in $C$. \end{pf} \begin{theorem} \label{symmetric} If $f \in {\cal F}$ and $f^\ast=-f$, then $\langle \alpha,\beta \rangle_f = \langle \beta,\alpha \rangle_f$ for all $\alpha,\beta \in \ker f$. \end{theorem} \begin{pf} By definition of the pairing, $\langle \;,\; \rangle_{-f} = -\langle \;,\; \rangle_f$, so \begin{eqnarray*} \langle \alpha,\beta \rangle_f & = & -\langle \beta,\alpha \rangle_{f^\ast} \quad \text{(by Proposition~\ref{backwards})} \\ & = & -\langle \beta,\alpha \rangle_{-f} \\ & = & \langle \beta,\alpha \rangle_f. \end{eqnarray*} \end{pf} \begin{rems} (I thank Noam Elkies for these.) The theory of additive polynomials shares much with the theory of differential operators, as pointed out by Ore~\cite{Or1}. In fact the pairing $\langle \;,\; \rangle_f$ can be obtained an analogue of a known pairing on kernels of differential operators. If $f$ is a differential operator and $f^\ast$ denotes its formal adjoint, then we have an identity $uf^\ast(v)-vf(u)=B'$ where $B$ is a bilinear form in $\{u,u',u'',\ldots,u^{(n-1)}\}$ and $\{v,v',v'',\ldots,v^{(n-1)}\}$. (See p.\ 124 in~\cite{ince}, where $B$ is called a ``bilinear concomitant.'') If $u \in \ker f$ and $v \in \ker f^\ast$, then $B(u,v)$ is a constant depending bilinearly on $u$ and $v$. Similarly, if $f \in {\cal F}$, then one can write $uf^\ast(v)-vf(u)=B^q-B$ where $B$ is a (potentially infinite) linear combination of terms of the form $u^{q^m} v^{q^n}$ ($m,n \in {\Bbb Z}$). One can check that $B$ converges to a continuous bilinear function on $C \times C$ by writing $B$ explicitly in terms of the coefficients of $f$, and using Proposition~\ref{convergence}. Clearly $B(\alpha,\beta) \in {{\Bbb F}_q}$ if $\alpha \in \ker f$ and $\beta \in \ker f^\ast$. \end{rems} \begin{prop} \label{ince} If $\alpha \in \ker f$ and $\beta \in \ker f^\ast$, then $B(\alpha,\beta)=\langle \alpha,\beta \rangle_f$. \end{prop} \begin{pf} Let $b_\alpha(v)=B(\alpha,v)$. If we set $u=\alpha$ in the defining equation for $B$, we have as functions of $v$, \begin{eqnarray*} \alpha \circ f^\ast & = & (\tau-1) \circ b_\alpha, \\ f \circ \alpha & = & b_\alpha^\ast \circ (\tau^{-1}-1) \\ & = & b_\alpha^\ast \circ \tau^{-1} \circ (1 -\tau). \end{eqnarray*} Comparing with equation~(\ref{def}) shows that $g_\alpha = b_\alpha^\ast \circ \tau^{-1}$, so \begin{eqnarray*} \langle \alpha,\beta \rangle_f & = & g_\alpha^\ast(\beta) \\ & = & (\tau \circ b_\alpha)(\beta) \\ & = & \tau(B(\alpha,\beta)) \\ & = & B(\alpha,\beta), \end{eqnarray*} since $B(\alpha,\beta) \in {{\Bbb F}_q}$. \end{pf} This alternative definition of the pairing could clearly be used to give new proofs of the properties of the pairing. For instance, Proposition~\ref{backwards} would be immediate, as would the first part of Theorem~\ref{skewsymmetric}. \section{$A$-module pairings for Drinfeld modules} \label{amodule} Let us retain the assumption that $A$ is the affine ring of a nonsingular projective curve over ${{\Bbb F}_q}$ minus a closed point $\infty$, and suppose we have an ${{\Bbb F}_q}$-algebra homomorphism $\iota: A \rightarrow C$. The kernel ${\frak p}$ of $\iota$ will be called the characteristic. Let $a \mapsto \phi_a$ and $a \mapsto \psi_a$ be ring homomorphisms from $A$ to ${\cal F}$. For example, $\phi$ and $\psi$ might be Drinfeld $A$-modules over $C$. We say that a nonzero $f \in {\cal F}$ is an {\em isogeny} from $\phi$ to $\psi$ if $\psi_a \circ f = f \circ \phi_a$ for all $a \in A$. If $f: \phi \rightarrow \psi$ is an isogeny, then $\ker f$ is an $A$-module via $\phi$ and $\ker f^\ast$ is an $A$-module via $\psi^\ast$. \begin{prop} \label{apairing} Let $a \mapsto \phi_a$ and $a \mapsto \psi_a$ be ring homomorphisms from $A$ to ${\cal F}$. Suppose $f \in {\cal F}$ is an isogeny from $\phi$ to $\psi$. Then for all $a \in A$, $\alpha \in \ker f$, $\beta \in \ker f^\ast$, we have $$\langle \phi_a(\alpha),\beta \rangle_f = \langle \alpha,\psi_a^\ast(\beta) \rangle_f,$$ and \begin{eqnarray*} [ \;\; , \;\; ]_f : \ker f \times \ker f^\ast & \rightarrow & \operatorname{Hom}_{{\Bbb F}_q}(A,{{\Bbb F}_q}) \\ \alpha \; , \; \beta \;\;\;\;\;\;\; & \mapsto & (a \mapsto \langle \phi_a(\alpha),\beta \rangle) \end{eqnarray*} is a pairing of $A$-modules which exhibits $\ker f^\ast$ as the Pontryagin dual of $\ker f$ as an $A$-module. Furthermore, if $\Psi:A \rightarrow {\cal F}$ is a third ring homomorphism, and $h:\Psi \rightarrow \phi$ is another isogeny, then we have the following two compatibility relations: if $\alpha \in \ker(f\circ h),\beta \in \ker f^\ast \subseteq \ker (f \circ h)^\ast$, then $$[ \alpha,\beta ]_{f \circ h} = [ h(\alpha),\beta ]_f.$$ Similarly, for all $\alpha \in \ker h \subseteq \ker(f \circ h), \beta \in \ker(f \circ h)^\ast$, $$[ \alpha,\beta ]_{f \circ h} = [ \alpha,f^\ast(\beta) ]_h.$$ \end{prop} \begin{pf} Given $a \in A$, $\alpha \in \ker f$, $\beta \in \ker f^\ast$, we have \begin{eqnarray*} \langle \phi_a(\alpha),\beta \rangle_f & = & \langle \alpha,\beta \rangle_{f \circ \phi_a} \;\;\; \text{(by Proposition~\ref{compatible})} \\ & = & \langle \alpha,\beta \rangle_{\psi_a \circ f} \\ & = & \langle \alpha,\psi_a^\ast(\beta) \rangle_f \;\;\; \text{(by Proposition~\ref{compatible} again}). \end{eqnarray*} Apply Lemma~\ref{showdual3} to get the duality property of $[\;\;,\;\;]_f$. The compatibility relations follow from those for $\langle \;\;,\;\; \rangle$ proved in Proposition~\ref{compatible}. \end{pf} \begin{cor} \label{weil} Let $\phi$ be a Drinfeld $A$-module over a field $L$. Fix a nonzero $a \in A$. Let $\phi[a]$ be the kernel of $\phi_a$ on $\overline{L}$, and similarly define $\phi^\ast[a]=\ker \phi_a^\ast$. There is a Galois-equivariant perfect pairing of finite $A$-modules $$[ \;\;,\;\; ]_a: \phi[a] \times \phi^\ast[a] \rightarrow \widehat{(A/a)} \subset \hat{A}.$$ (Here $\operatorname{Gal}(L^{\text{sep}}/L)$ acts trivially on $\hat{A}$ and $\widehat{(A/a)}$.) If we also have a nonzero $b \in A$, then for any $\alpha \in \phi[ab]$, $\beta \in \psi^\ast[a]$ $$[ \alpha,\beta ]_{ab} = [\phi_b(\alpha),\beta]_a.$$ Similarly, for $\alpha \in \phi[a]$, $\beta \in \psi^\ast[ab]$, $$[ \alpha,\beta ]_{ab} = [\alpha,\psi^\ast_b(\beta)]_a.$$ \end{cor} The properties above of $[\;\;,\;\;]_a$ should remind one of the Weil pairing for abelian varieties. Another proof of the Galois-equivariant duality between $\phi[a]$ and $\phi^\ast[a]$ was discovered independently by Taguchi. (See the appendix of~\cite{gossadjoint}.) As another application of our pairings, we can describe the kernel of the adjoint of the exponential function associated with a Drinfeld module. \begin{cor} \label{estar} Let $C$ be the completion of the algebraic closure of $K_\infty$ with $|\;|_\infty$, and let $\phi$ be a Drinfeld $A$-module over $C$. Let $e(z) \in C[[z]]$ be the associated exponential function. Then there is a natural pairing $$\ker e \times \ker e^\ast \rightarrow \hat{A}$$ which exhibits $\ker e^\ast$ as the Pontryagin dual of $\ker e$ as an $A$-module. If the lattice $\ker e$ is isomorphic to the direct sum of fractional ideals $I_1 \oplus \cdots \oplus I_r$, then $\ker e^\ast$ is isomorphic to $K_\infty/(I_1^{-1}J) \oplus \cdots \oplus K_\infty/(I_r^{-1}J)$ as a topological $A$-module. (Here $J$ is as in Theorem~\ref{fractionaldual}.) \end{cor} \begin{pf} Simply note that $e$ is an isogeny from $C$ with the standard $A$-module structure to $\phi$, and apply Proposition~\ref{apairing}. The last assertion follows from Theorem~\ref{fractionaldual}. \end{pf} \begin{prop} Let $\phi$ and $e$ be as in the previous corollary. Then $\ker e^\ast$ is the closure of $\bigcup_{a \in A} \phi^\ast[a]$. \end{prop} \begin{pf} Taking adjoints of $$e \circ a = \phi_a \circ e$$ yields $$a \circ e^\ast = e^\ast \circ \phi^\ast_a,$$ from which it is clear that $\phi^\ast[a] \subset \ker e^\ast$. Corollary~\ref{estar} says that $\ker e^\ast$ as a topological $A$-module via $\phi^\ast$ is isomorphic to $K_\infty/C_1 \oplus \cdots \oplus K_\infty/C_r$, for some fractional $A$-ideals $C_1,\ldots,C_r$. The torsion submodule in this module is $K/C_1 \oplus \cdots \oplus K/C_r$, which is dense, so the result follows. \end{pf} \section{Tate module pairings} Because of the compatibility relations, we can use our pairings to construct a pairing between Tate modules, just as the Weil pairing gives rise to a pairing between Tate modules. Let ${\frak q}$ be a nonzero prime of $A$ different from the characteristic ${\frak p}$ of the Drinfeld module. Let $K$ be the fraction field of $A$, and let $A_{\frak q}$, $K_{\frak q}$ be the completions at the prime ${\frak q}$. The Tate module of a Drinfeld $A$-module over $L$ is $$T_{\frak q}(\phi) = \operatorname{Hom}_A(K_{\frak q}/A_{\frak q},\phi(\overline{L}))$$ where $\phi(\overline{L})$ denotes the additive group of $\overline{L}$ with the $A$-module structure given by $\phi$. Similarly we define the Tate module of $\phi^\ast$ as $$T_{\frak q}(\phi^\ast) = \operatorname{Hom}_A(K_{\frak q}/A_{\frak q},\phi^\ast(\overline{L})).$$ These are both free $A_{\frak q}$-modules of rank equal to the rank of the Drinfeld module. \begin{prop} \label{tate} There is a Galois-equivariant continuous perfect pairing of $A_{\frak q}$-modules $$[\;\;,\;\;]_{\frak q} : T_{\frak q}(\phi) \times T_{\frak q}(\phi^\ast) \rightarrow \operatorname{Hom}_{{\Bbb F}_q}(K_{\frak q}/A_{\frak q},{{\Bbb F}_q}) \cong A_{\frak q}$$. \end{prop} \begin{pf} Given $\alpha \in T_{\frak q}(\phi)$, $\beta \in T_{\frak q}(\phi^\ast)$, define $[\alpha,\beta]_{\frak q} \in \operatorname{Hom}_{{\Bbb F}_q}(K_{\frak q}/A_{\frak q},{{\Bbb F}_q})$ by $$[\alpha,\beta]_{\frak q}(b)=[\alpha(b),\beta(a^{-1})]_a$$ where $a \in A$ kills $b$ and generates an ideal power of ${\frak q}$. As is well known, $A$ is a Dedekind domain with finite class number, so some power ${\frak q}^h$ of ${\frak q}$ is principal, so it is always possible to find such an $a$, given $b$. The definition is independent of the choice of $a$, by the last compatibility relation in Corollary~\ref{weil}. If we fix $a \in A$ generating a power of ${\frak q}$, we can also obtain the pairing as the inverse limit of the pairings $$[\;\;,\;\;]_{a^n} : \phi[a^n] \times \phi^\ast[a^n] \longrightarrow \operatorname{Hom}_{{\Bbb F}_q}(A/a^n,{{\Bbb F}_q}) = \operatorname{Hom}_{{\Bbb F}_q}(a^{-n}A/A,{{\Bbb F}_q}),$$ with respect to the maps \begin{eqnarray*} \phi[a^{n+1}] & \stackrel{a}{\longrightarrow} & \phi[a^n] \\ \phi^\ast[a^{n+1}] & \stackrel{a}{\longrightarrow} & \phi^\ast[a^n] \\ \operatorname{Hom}_{{\Bbb F}_q}(a^{-(n+1)}A/A,{{\Bbb F}_q}) & \stackrel{\text{res}}{\longrightarrow} & \operatorname{Hom}_{{\Bbb F}_q}(a^{-n}A/A,{{\Bbb F}_q}). \end{eqnarray*} The Galois-equivariance, continuity, and perfectness of the pairing then follow from the properties of the pairings $[\;\;,\;\;]_{a^n}$. Finally, $\operatorname{Hom}_{{\Bbb F}_q}(K_{\frak q}/A_{\frak q},{{\Bbb F}_q})$ is (non-canonically) isomorphic to $A_{\frak q}$ as an $A_{\frak q}$-module, by Theorem~\ref{aqdual}. \end{pf} As pointed out by Goss, this duality of Tate modules has the following corollary. \begin{cor} \label{semisimple} Suppose the Drinfeld module $\phi$ is defined over a finite extension $L$ of $K$. Then $$T_{\frak q}(\phi^\ast) \otimes_{A_{\frak q}} K_{\frak q}$$ is a semi-simple $K_{\frak q}[\operatorname{Gal}(L^{\text{sep}}/L)]$-module. \end{cor} \begin{pf} Combine the previous theorem with the main theorem in~\cite{taguchi}. \end{pf} \section{Questions} \label{questions} Our results show that in two cohomological realizations (the period lattice and Tate module), the cohomology of the adjoint of a Drinfeld module is dual to the cohomology of the original Drinfeld module. There is a third cohomological realization of a Drinfeld module, namely the de Rham cohomology~\cite{gekeler} developed by Anderson, Deligne, Gekeler, and Yu. \begin{question} Is it possible to give a reasonable definition for the de Rham cohomology of the adjoint of a Drinfeld module and prove that it is dual in some sense to the de Rham cohomology of the original Drinfeld module? \end{question} It would also be nice to generalize the applications of this paper to include $t$-modules, the higher dimensional analogues of Drinfeld modules. \begin{question} Can one prove results similar to those in this paper for fractional power series in more than one variable? \end{question} A reasonable approach is to work with the ring ${{\Bbb M}_d}({\cal F})$ of $d$-by-$d$ matrices with coefficients in ${\cal F}$. Elements of this ring act on $C^d$ in an obvious way. Moreover, there is a multiplication-reversing adjoint map ${{\Bbb M}_d}({\cal F}) \rightarrow {{\Bbb M}_d}({\cal F})$, which maps a matrix $A=(a_{ij})$ to $A^\ast=(a_{ji}^\ast)$, i.e., the transpose of the matrix obtained by taking the adjoint of each entry. Hence, for example, it is possible to define the adjoint of a $t$-module \begin{eqnarray*} \phi: {{\Bbb F}_q}[t] & \rightarrow & {{\Bbb M}_d}(C\{\tau\}) \subset {{\Bbb M}_d}({\cal F}) \\ a & \mapsto & \phi_a \end{eqnarray*} as \begin{eqnarray*} \phi^\ast: {{\Bbb F}_q}[t] & \rightarrow & {{\Bbb M}_d}({\cal F}) \\ a & \mapsto & \phi_a^\ast. \end{eqnarray*} For each $A = (a_{ij}) \in {{\Bbb M}_d}({\cal F})$ it is also possible to define a natural bilinear pairing $$\langle \;,\; \rangle_A : \ker A \times \ker A^\ast \rightarrow {{\Bbb F}_q}$$ by following the differential operator analogy. Let $\cdot$ denote the standard inner product on $C^d$. As in the remarks at the end of Section~\ref{adjointpd}, for each $f \in {\cal F}$, let $B_f(u,v)$ be the bilinear function of $u,v \in C$ such that $$u f^\ast(v) - v f(u) = B_f^q - B_f.$$ Then, for $u,v \in C^d$, the bilinear function $$B_A(u,v) = \sum_{i,j} B_{a_{ij}}(u_i,v_j)$$ satisfies $$u \cdot A^\ast(v) - v \cdot A(u) = B_A^q-B_A,$$ and we may define $$\langle u,v \rangle_A = B_A(u,v).$$ Thus for instance, one has a pairing between the $a$-torsion of a $t$-module and the $a$-torsion of its adjoint. \section*{Acknowledgements} Many thanks go to David Goss, who introduced me to adjoints of polynomials and Drinfeld modules, and asked some of the questions answered by this paper. I thank T.\ Y.\ Lam, Hendrik Lenstra, and Ken Ribet for helping me find references for Newton polygons. Thanks also to Noam Elkies for sharing his preprint with me, and to Michael Rosen for suggesting references of use in proving Theorem~\ref{fractionaldual}. Finally, I thank the referee for many intelligent and helpful suggestions.
\setcounter{equation}{0}\section@ori{\setcounter{equation}{0}\setcounter{equation}{0}\section@ori@ori} } \makeatother \makeatletter \newlength{\lr@listTW} \newlength{\lr@listLW} \def\lr@list#1{\def\lr@listDL{#1}% \settowidth{\lr@listLW}{\lr@listDL}% \lr@listTW=\linewidth% \addtolength{\lr@listTW}{-\lr@listLW}% \hbox to\lr@listLW{}\begin{minipage}[t]{\lr@listTW } \def\lrlist{% \def\lr@listPAR{\noindent}% \def\lr@listPL[##1] {% \lr@listPAR\raisebox{0pt}[0pt][0pt]{\llap{% \parbox[t]{\lr@listLW}{##1}}}% \def\lr@listPAR{\par\noindent}}% \def\item{\@ifnextchar [{\lr@listPL}{\lr@listPDL}}% \def\lr@listPDL{% \lr@listPAR\raisebox{0pt}[0pt][0pt]{\llap{% \parbox[t]{\lr@listLW}{\lr@listDL}}}% \def\lr@listPAR{\par\noindent}}% \lr@list}% \def\end{minipage}{\end{minipage}} \makeatother \def\figurename{Fig.} \setlength{\textfloatsep}{0.4\baselineskip plus 2pt minus 4pt} \ifx\LRcaptionwidth\UNDEFINE \newlength{\LRcaptionwidth}\else\fi \LRcaptionwidth=\textwidth \addtolength{\LRcaptionwidth}{-2\parindent} \begin{document} \hbadness=10000 \ifx\hyperref\undefined\else\errmessage{HYPER DISABLED}\fi \setcounter{page}{1} \thispagestyle{empty} \begin{samepage} \title{% {\normalsize\sf% {\sl August 1995\/}\hfill \begin{tabular}[t]{l} GSI-Preprint-95-46 \\ hep-ph/9508318 \end{tabular}}\\*[5ex] {\bf Photon Intensity Interferometry\\ for Expanding Sources} } \author{Leonid V. Razumov% \thanks{E.Mail: \sf\,\, razumov@clri6a.gsi.de\ \ (URL$\Rightarrow$ http://www.gsi.de/groups/the/razumov.html)}\,\,, \quad Hans Feldmeier\setcounter{footnote}{3}% \thanks{E.Mail: \sf\,\, <EMAIL>} } \date{Gesellschaft f\"ur Schwerionenforschung, Postfach 110552,\\ D-64220 Darmstadt, Germany} \maketitle \setcounter{footnote}{0} \thispagestyle{empty} \begin{center} {\large\bf Abstract}\\[2ex] \end{center} {\narrower\narrower\par\noindent \small Using Quantum Field Theory we derive a general formula for the double inclusive spectra of photons radiated by a system in local equilibrium. The derived expression differs significantly from the one mostly used up to now in photon intensity interferometry of heavy--ion collisions. We present a covariant expression for double inclusive spectra adapted for usage in numerical simulations. Application to a schematic model with a Bj\o rken type expansion gives strong evidence for the need of reinvestigating photon--photon correlations for expanding sources. \par} \vskip 15mm \noindent \centerline{\large\sl Submitted to Physics Letters B\/.} \end{samepage} \newpage \noindent Double inclusive spectra of hard photons radiated at the early stage of heavy--ion collisions have become an important subject of theoretical \cite{Neuh86,SriKap93a,TimPlumRazWein94} and experimental \cite{GANIL94} investigations. Since these photons originate from very energetic collisions and interact very little with the matter they are traversing, they may carry signatures of the hot and dense phase of the matter, in particular of the quark--gluon plasma. Except for the work of D.~Neuhauser \cite{Neuh86} in which he treated a static source, usually one is analyzing the correlations in two--photon coincidence measurements with expressions based on the corresponding ones for scalar bosons \cite{Pratt84} augmented with a degeneracy factor which is supposed to take care of the spin structure of the photons. The aim of this work is to derive in a quantum--field treatment covariant expressions for the two--photon density and herewith the corresponding correlation function. In particular we pay attention to charge conservation and to the fact that massless photons have only two helicity eigenstates. Due to these properties averaging over all spin directions results in a reduction factor of one--half compared to the scalar case only if the momenta of the two photons are parallel. For a finite angle, however, the interference is reduced, so that the deduced size of the source is smaller than the one obtained from the correlation function for scalar bosons. For Bj\o{}rken type hydrodynamics we estimate a difference in the apparent source size of about 30\% in rapidity. Therefore, we conclude that the ``pion--inspired'' formulas for the Bose--Einstein correlation function is not applicable for photon intensity interferometry studies, especially if one deals with an expanding source. The finally measured number of photons per invariant momentum bin with polarization $\,\lambda\,$ is given by the density matrix of the final state $\rho_f$ which can be expressed in terms of the initial density matrix $\rho_i$ and the $S$--matrix using the relation \mbox{$\rho_f = S \rho_i S^{\dag}$}. \begin{equation} \label{P1:def} P_1^{\lambda}(\mbold{k})\ \equiv\ k^0 \frac{d N^{\lambda}}{d^3\mbold{k}} = Tr\{ \rho_f c_{\lambda}^{\dag}(\mbold{k}) c_{\lambda}(\mbold{k})\} = Tr\{ \rho_i S^{\dag} c_{\lambda}^{\dag}(\mbold{k}) c_{\lambda}(\mbold{k}) S \} \end{equation} The same consideration as above leads to the expression for the number of photon pairs \ba \label{P2:def} P_2^{\lambda_1\;\lambda_2}(\mbold{k}_1,\mbold{k}_2)\ &\equiv&\ k^0_1k^0_2 \frac{d^2 N^{\lambda_1 \lambda_2}}% {d^3 \mbold{k}_1 d^3 \mbold{k}_2} = Tr\{\rho_f c_{\lambda_2}^{\dag}(\mbold{k}_2) c_{\lambda_1}^{\dag}(\mbold{k}_1) c_{\lambda_1}(\mbold{k}_1) c_{\lambda_2}(\mbold{k}_2)\}\ = \nonumber\\ &=&\ Tr\{\rho_iS^{\dag} c_{\lambda_2}^{\dag}(\mbold{k}_2) c_{\lambda_1}^{\dag}(\mbold{k}_1) c_{\lambda_1}(\mbold{k}_1) c_{\lambda_2}(\mbold{k}_2) S\}\;\ . \quad \end{eqnarray} If we separate the interaction part of the Lagrangian into $L^{int}_s(x)$ for strong interactions and $J^{\mu}(x)A_{\mu}(x)$ for the coupling to the electromagnetic field the \mbox{$S$--matix} can be written in the interaction picture by the time--ordered exponential \cite{BjorkenDrell65} \begin{equation} \label {S:def} S\ =\ {\cal T}\exp\left\{i\int d^4x \Big( L^{int}_s(x) + J^{\mu}(x)A_{\mu}(x)\Big) \right\}\;\ .\quad \end{equation} To proceed further we assume \cite{Machl89} that the initial state does not contain hard photons, which means \begin{equation} \label{rhoi:def} c_{\lambda}(\mbold{k})\rho_i\ =\ \rho_i c^{\dag}_{\lambda}(\mbold{k}) \ =\ 0 \;\ .\quad \end{equation} The condition (\ref{rhoi:def}) together with the identities \ba \Big[ c_{\lambda}(\mbold{k})\,,\, S \Big] \ &=&\ i {\cal T} \Big(J_{\lambda}(k) S \Big) \quad, \label{commut:cS} \\\relax \Big[ S^{\dag}\,,\, c^{\dag}_{\lambda}(\mbold{k}) \Big] \ &=&\ -i \tilde{\cal T} \Big(S^{\dag} J^{\,\,\dag}_{\lambda}(k) \Big) \quad, \nonumber \end{eqnarray} where $k^{\mu}=(|\mbold{k}|,\mbold{k})$ and $J_{\lambda}(k)$ is the Fourier transform of the transverse current operator given by \ba J_{\lambda}(k)\ &\equiv&\ \epsilon^{\!\!^{\scriptstyle *}\lambda}_{\mu}(k) \!\int\! d^4\!x\,{\rm e}^{ixk} J^{\mu}(x) \label{J_lambda:def} \\ J^{\,\,\dag}_{\lambda}(k)\ &\equiv&\ \epsilon^{\lambda}_{\mu}(k) \!\int\! d^4\!x\,{\rm e}^{-ixk} J^{\mu}(x) \quad, \nonumber \end{eqnarray} allow us to write one-- and two--body densities via the chronological (${\cal T}$) and antichronological ($\tilde{\cal T}$) products of the current operators: \ba P_1^{\lambda}(\mbold{k})\ &=&\ Tr \bigg\{ \rho_i % \tilde{\cal T} \Big( S^{\dag} J^{\,\,\dag}_{\lambda}(k) \Big) {\cal T}\Big( J_{\lambda}(k) S \Big) \bigg\} \label{P1:S} \\ P_2^{\lambda_1 \lambda_2}(\mbold{k}_1, \mbold{k}_2)\ &=&\ Tr \bigg\{ \rho_i \tilde{\cal T} \Big( S^{\dag}% J^{\,\,\dag}_{\lambda_2}(k_2) J^{\,\,\dag}_{\lambda_1}(k_1) \Big) {\cal T} \Big( J_{\lambda_1}(k_1) J_{\lambda_2}(k_2) S \Big)\bigg\} \quad . \label{P2:S} \end{eqnarray} All currents in the expressions written above are operators in the {\sl interaction} representation. Utilizing properties of time ordering and of the $S$--matrix we can write the one-- and two--photon spectra in terms of the current operators in the {\sl Heisenberg} representation ($\hat{J})$ \cite{RazWein95} as \ba P_1^{\lambda}(\mbold{k})\ &=&\ Tr \bigg\{ \rho_i % \hat{J}^{\,\,\dag}_{\lambda}(k) \hat{J}_{\lambda}(k) \bigg\} \label{P1:H} \\ P_2^{\lambda_1 \lambda_2}(\mbold{k}_1, \mbold{k}_2)\ &=&\ Tr \bigg\{ \rho_i \tilde{\cal T} \Big( % \hat{J}^{\,\,\dag}_{\lambda_2}(k_2) \hat{J}^{\,\,\dag}_{\lambda_1}(k_1) \Big) {\cal T} \Big( \hat{J}_{\lambda_1}(k_1) \hat{J}_{\lambda_2}(k_2) \Big)\bigg\} \quad. \label{P2:H} \end{eqnarray} Up to now we have made no approximation to tackle the unsolvable many--body problem. Expressions (\ref{P1:H}) and (\ref{P2:H}) are still exact but they are recast into a form which is better suited for approximations than the original definitions (\ref{P1:def}) and (\ref{P2:def}). Formally they contain only operators of the strongly interacting system. In the following we apply eqs. (\ref{P1:H}) and (\ref{P2:H}) to a highly excited system which due to short--ranged hard collisions of the charged constituents is radiating energetic photons. The collisions cause locally rapid changes in the electric charge current density $J^{\mu}(x)\,$, which result in non--zero Fourier components in $J^{\mu}(k)\,$ for large $|\mbold{k}|\,$. The energetic photons are weakly interacting so that, once they are created, they are assumed to leave the hadron system without further interactions. Therefore, for calculating the final photon distributions one needs a local production rate which is then integrated over space and time. Eq. (\ref{P1:H}) is of this type. In coordinate space it reads \begin{equation} \label{P1:xdx} P^{\lambda}_1(\mbold{k})= \epsilon^{\!\!^{\scriptstyle *}\lambda}_{\mu_1}(k) \epsilon^{\lambda}_{\mu_2}(k)\!\! \int\!\!d^4\!\bar{x}\!\!\int\!\!d^4\!\Delta{}x\, {\rm e}^{-ik\Delta x} \,Tr\Big\{\rho_i \hat{J}^{\mu_1}(\bar{x}+\frac{\Delta x}{2}) \hat{J}^{\mu_2}(\bar{x}-\frac{\Delta x}{2}) \Big\} \end{equation} where the local rate is given by the integral over $\Delta x\,$ of the current--current correlator. The following approximations, which will lead to an expression for the correlator adapted for usage in numerical simulations for heavy--ion collisions, are based upon three physical assumptions: \begin{LRdescription}{$(iii)$\hbox{\,\ }} \item[\hfill$(i)$\hbox{\,\ }] The hadronic correlations in the system are of short range in space--time. The typical correlation length is about the mean free path $\lambda_s$ of the strongly interacting particles. Therefore, \benn <\!A(x) B(y)\!> \approx <\!A(x)\!> <\!B(y)\!> \qquad \mbox{\rm if}\;\ |\mbold{x} - \mbold{y}| \;\gap\; \lambda_s \;\ \mbox{\rm or}\;\ |x^0 - y^0| \;\gap\; \lambda_s\ . \eenn \item[\hfill$(ii)$\hbox{\,\ }] The hadronic mean free path $\lambda_s$ is much less than the characteristic size of variations of the macroscopic variables in space and time, denoted by $L\,$. \item[\hfill$(iii)$\hbox{\,\ }] Only hard photons with high momenta $|\mbold{k}|$ are considered $\,(\,|\mbold{k}| \lambda_s \;\gap\; 1\,)\,$, for which the slowly varying collective current $<\!\!\hat{J}^{\mu}(x)\!\!>$ (proportional to the collective 4--velocity field $u^{\mu}(x)\,$) does not contribute, i.e. $<\!\!\hat{J}^{\mu}(k)\!\!> \approx 0 $ for high enough $|\mbold{k}|\,$. \end{LRdescription} For instance, applying these assumptions to (\ref{P1:xdx}) means that the integral in $\Delta x$ effectively spreads over the 4--volume of the order $\lambda_s^4$ and the single inclusive spectrum $P_1(\mbold{k})\,$ is therefore proportional to $L^4 \lambda^4_s\;$. Inside the trace on the right--hand side of (\ref{P1:xdx}) the mean time $\bar{x}^0\,$ can be moved over to the statistical operator such that \begin{equation} \label{JJ:trick} Tr\Big\{\rho_i \hat{J}^{\mu_1}(\bar{x}+\frac{\Delta x}{2}) \hat{J}^{\mu_2}(\bar{x}-\frac{\Delta x}{2}) \Big\} = Tr\Big\{\rho(\bar{x}^0) \hat{J}^{\mu_1}(\tilde{x}+\frac{\Delta x}{2}) \hat{J}^{\mu_2}(\tilde{x}-\frac{\Delta x}{2}) \Big\}\ ,\qquad \end{equation} where $\tilde{x}=(0,\mbold{x})$ and $\rho(\bar{x}^0)={\rm e}^{-iH\bar{x}^0} \rho_i {\rm e}^{iH\bar{x}^0}$ is the solution of the Liouville equation for the density matrix. For the case of local equilibrium, where the time--dependence enters the density matrix $\,\rho(\bar{x}^0)\,$ only through the thermodynamic quantities, like for example temperature $T(\bar{x}^0,\mbold{x})\,$, 4--velocity $\,u^{\mu}(\bar{x}^0,\mbold{x})\,$ etc., the identity (\ref{JJ:trick}) is a starting point to express the current--current correlator in terms of thermodynamic variables. The double inclusive spectrum is given by the four--point correlator (\ref{P2:H}) and like any other Green function can be written in terms of connected parts as \begin{samepage} \ba P_2^{\lambda_1 \lambda_2}(\mbold{k}_1, \mbold{k}_2) =&& P_1^{\lambda_1}(\mbold{k}_1) P_1^{\lambda_2}(\mbold{k}_2) + |\!\!<\!\!\hat{J}^{\dag}_{\lambda_1}(k_1) \hat{J}_{\lambda_2}(k_2)\!\!>\!\!|^2 + |\!\!<\!\!{\cal T} \Big( \hat{J}_{\lambda_1}(k_1) \hat{J}_{\lambda_2}(k_2) \Big)\!\!>\!\!|^2 \qquad \nonumber \\ && \;+\; \ll\!\! \tilde{\cal T} \Big( \hat{J}^{\,\,\dag}_{\lambda_2}(k_2) \hat{J}^{\,\,\dag}_{\lambda_1}(k_1) \Big) {\cal T} \Big( \hat{J}_{\lambda_1}(k_1) \hat{J}_{\lambda_2}(k_2) \Big) \!\!\gg\ .\qquad \label{P2:JJ} \end{eqnarray} \end{samepage} All terms in (\ref{P2:JJ}) involve four space--time integrations. In coordinate space a connected correlator disappears if the distance between any pair of its space--time points exceeds the correlation length. The connected four--point correlator $\,\ll\!\!\cdots\!\!\gg\,$ (last term on the right--hand side of (\ref{P2:JJ})\,) is only non--zero if the relative distance of all its space--time points are within $\,\lambda_s\,$. Therefore, this term is proportional to $\,L^4 (\lambda^4_s)^3\,$ and hence, due to condition $\,(ii)\,$, is smaller by factor of $\,(\lambda_s/L)^4\,$ when compared to all other terms which are proportional to $\,(L^4)^2(\lambda^4_s)^2\,$. The third term on the r.h.s of (\ref{P2:JJ}), $\,|\!\!<\!\!{\cal T} \Big( \hat{J}_{\lambda_1}(k_1) \hat{J}_{\lambda_2}(k_2) \Big)\!\!>\!\!|^2\,$, contains an integral $\,\int\!d^4\!\bar{x}\, \exp(i(|\mbold{k}_1|+|\mbold{k}_2|)\bar{x}^0)\,\cdots \;$ in which the energies of the two photons add up in the phase factor. This implies that only momenta which fulfill the condition $\,(|\mbold{k}_1|+|\mbold{k}_2|)\cdot L \;\lap\; 1 \,$ contribute appreciably to this term. These momenta are much smaller than required by conditions $\,(ii)\,$ and $\,(iii)\,$ and hence we shall drop this term. After these approximations based on the statistical properties of the emitting source the one-- and two--photon spectra (\ref{P1:H}) and (\ref{P2:H}) are given by the current--current correlator $\,<\!\!\hat{J}^{\mu_1 \dag}(k_1)\hat{J}^{\mu_2}(k_2)\!\!>\,$ only. Because the electrical current is conserved $(\partial_{\mu}\hat{J}^\mu(x)=0)$ the correlator must be transverse ``from both sides'': \begin{equation} \label{kJJ=0} k_{1 \mu_1}<\!\!\hat{J}^{\mu_1 \dag}(k_1)\hat{J}^{\mu_2}(k_2)\!\!> = <\!\!\hat{J}^{\mu_1 \dag}(k_1)\hat{J}^{\mu_2}(k_2)\!\!>k_{2 \mu_2} = 0 \;\ .\qquad\end{equation} In analogy to (\ref{P1:xdx}) the correlator can be written as \ba <\!\!\hat{J}^{\mu_1 \dag}(k_1)\hat{J}^{\mu_2}(k_2)\!\!> = \sum\limits_{n}&&{\rm e}^{-i(k_1-k_2)\bar{x}_n}\!\! \int\limits_{\Omega_n}\!\!d^4\!\bar{x} {\rm e}^{-i(k_1-k_2)(\bar{x}-\bar{x}_n)}\!\!\int\!\!d^4\!\Delta x {\rm e}^{-i\Delta x (k_1+k_2)/2}\, \hbox{\qquad}\nonumber\\ \times&&Tr\Big\{\rho_i \hat{J}^{\mu_1}(\bar{x}+\frac{\Delta x}{2}) \hat{J}^{\mu_2}(\bar{x}-\frac{\Delta x}{2})\Big\}\;\ , \qquad \label{JJ:xdx} \end{eqnarray} where the whole space--time region occupied by the source is divided into cells with 4--volume $\,\Omega_n\,$ located at mean--coordinates $\,\bar{x}_n\,$. Under the assumptions $\,(i)-(iii)\,$ the size of the space-time cells should be chosen to be about $\lambda_s$ so that the radiation of hard photons from different cells can be considered as entirely independent processes. This implies that the currents are conserved for each cell separately and that the contribution of the cells to the total current correlator is additive. Therefore, we parametrize (\ref{JJ:xdx}) as \begin{equation} \label{JJ:W} <\!\!\hat{J}^{\mu_1 \dag}(k_1)\hat{J}^{\mu_2}(k_2)\!\!>\ \stackrel{ansatz}{=\!=\!=} \ \sum\limits_{n} \Omega_n{\rm e}^{-i\bar{x}_n(k_1 - k_2)} Q^{\mu_1\mu_2}(k_1,k_2|\bar{x}_n) w(k_1,k_2|\bar{x}_n)\;\ ,\quad \end{equation} where the tensor $Q^{\mu_1 \mu_2}$ carries the tensorial structure of the current--current correlator and the function $w$ describes the strength. The $Q$--tensor is explicitly transverse \begin{equation} \label{kQ=0} \,k_{1\mu_1}Q^{\mu_1\mu_2}(k_1,k_2|x) = Q^{\mu_1\mu_2}(k_1,k_2|x)k_{2\mu_2}\nolinebreak=\nolinebreak{}0\, \end{equation} and, therefore, ensures current conservations for every cell. The tensor $Q^{\mu_1\mu_2}$ and the function $w$ are hermitian in the sense that $Q^{\mu_1\mu_2}(k_1,k_2|x)^{\mbox{\boldmath$*$}} =Q^{\mu_2\mu_1}(k_2,k_1|x)$ and $w(k_1,k_2|x)^{\mbox{\boldmath$*$}} =w(k_2,k_1|x)$. It is convenient to normalize the $Q$--tensor by $\,Q^{\mu}_{\mu}(k,k|x) \equiv -2\,$. From now on we will write an integral instead of the sum in (\ref{JJ:W}). Altogether single and double inclusive photon spectra (no polarization measured) can be expressed with the help of (\ref{JJ:W}) as follows% \footnote{We take the sum over photon polarisations by means of \benn \!\sum\limits^2_{(\lambda=1)}\!% \epsilon^{\!\!^{\scriptstyle *}\lambda}_{\mu}(k)% \epsilon^{\lambda}_{\nu}(k) = - g_{\mu\nu} - k_{\mu}k_{\nu}/(sk)^2 + (k_{\mu}s_{\nu} + s_{\mu}k_{\nu})/(sk) \eenn ($\,s^{\mu}\,$ is some reference 4-vector $\,s^2=1\,$), where due to the transversality condition (\ref{kQ=0}) only the first term on the right--hand side $\,(-g_{\mu\nu})\,$ really contributes to (\ref{P1:W}), (\ref{P2:W}). }: \ba P_1(\mbold{k})\ &\equiv\ & \!\sum^2_{(\lambda=1)} P_1^{\lambda}(\mbold{k}) = 2 \!\int\! d^4\!x\, w(k,k|x) \label{P1:W} \\ P_2(\mbold{k}_1,\mbold{k}_2) &\equiv& \!\sum^2_{(\lambda_1,\lambda_2=1)}\!\! P_2^{\lambda_1 \lambda_2}(\mbold{k}_1,\mbold{k}_2) = P_1(\mbold{k}_1) P_1(\mbold{k}_2) \label{P2:W} \\ &+& \!\int\!d^4\!x d^4\!y \, {\rm e}^{-i(x-y)(k_1-k_2)} w(k_1,k_2|x) w(k_2,k_1|y) Q^{\mu_1\mu_2}(k_1,k_2|x) Q_{\mu_2\mu_1}(k_2,k_1|y) \quad \nonumber \end{eqnarray} To find an explicit form of the $Q$--tensor we assume that the radiating system is in local equilibrium, that is, the radiation from an elementary cell is isotropic in its local rest frame and therefore $Q^{i_1i_2}=\delta^{i_1i_2}$ (where $\,i_1,i_2=1,2,3$). The other components of the $Q$--tensor are then uniquely determined by the transversality condition (\ref{kQ=0}) as $\,Q^{00}=\mbold{k}_1\mbold{k}_2/k^0_1k^0_2\,$, $\,Q^{i_10}=k^{i_1}_2/k^0_2\,$, $\,Q^{0i_2}=k^{i_2}_1/k^0_1\,$. In local thermal equilibrium the only other 4--vector entering is the 4--velocity $u^{\mu}(x)$ of the emitting cell which is located at point $x\,$. Therefore, the only hermitian and transversal $Q$--tensor which can be constructed from $k^{\mu}_1\,,\; k_2^{\mu}\,\;$ and $u^{\mu}(x)\,$, and which in the rest frame of a cell (\,$u^{\mu}=(1,\mbox{\bf 0})$\,) reproduces the isotropic form given above, reads as follows: \begin{equation} \label{Q:leq} Q^{\mu_1\mu_2}(k_1,k_2|x) = -g^{\mu_1\mu_2} - u^{\mu_1}(x)u^{\mu_2}(x)\frac{(k_1k_2)}{(k_1u(x))(k_2u(x))} + \frac{u^{\mu_1}(x)k_1^{\mu_2}}{(u(x)k_1)} + \frac{k_2^{\mu_1}u^{\mu_2}(x)}{(u(x)k_2)}\ \end{equation} In eq.~(\ref{Q:leq}) we do not include the term proportional to $\,k_1^{\mu_1}k_2^{\mu_2}\,$ because it does not contribute to observables and can be removed by an appropriate gauge transformation. After defining the $Q$--tensor all dynamic information on the photon production is contained in the function $w(k_1,k_2|x)$. Inserting (\ref{Q:leq}) into (\ref{P2:W}) one gets \begin{equation} \label{P2:leq} \!\!\!% P_2(\mbold{k}_1,\mbold{k}_2) \!=\! P_1(\mbold{k}_1) P_1(\mbold{k}_2) \!+\!\!\int\!d^4\!x d^4\!y \cos(\Delta{}x\Delta{}k) R(k_1,k_2|x,y) w(k_1,k_2|x) w(k_2,k_1|y)\ \end{equation} where we introduce the abbreviation \ba &&R(k_1,k_2|x,y) \equiv Q^{\mu_1 \mu_2}(k_1,k_2|x)Q^{\mu_2 \mu_1}(k_2,k_1|y) = \frac{\big(u(x)k_1\big)\big(u(y)k_2\big) - 2\big(u(x)u(y)\big)\big(k_1k_2\big)}{% \big(u(x)k_2\big)\big(u(y)k_1\big)} \qquad \nonumber \\ &&+\ \frac{\big(k_1k_2\big)}{\big(k_1u(x)\big)\big(u(x)k_2\big)} \,+\, \frac{1}{2}\frac{\big(u(x)u(y)\big)^2\big(k_1k_2\big)^2}{% \big(k_1u(x)\big)\big(u(x)k_2\big)\big(k_1u(y)\big)\big(u(y)k_2\big)} \ +\ \Big(x \Leftrightarrow y \Big)\;\ . \quad \label{R:leq} \end{eqnarray} The $R$--function reflects the fact that photons are massless spin--1 particles and in this point expression (\ref{P2:leq}) differs significantly from the one obtained in \cite{Pratt84} for pions. The latter has been used for photons of one single polarization in \cite{SriKap93a,TimPlumRazWein94,SriKap93,SriKap94}. Since the experimental measurements will involve an averaging over polarizations, a comparison with data requires the new result (\ref{P2:leq}). Being very sensitive to the 4--velocity field $\,u^{\mu}(x)\,$ the $R$--function can change drastically the interference of two photons for a relativistically expanding source. But even for a static source the polarization properties of photons are very important. To illustrate this let us consider two photons with momentum $\,\mbold{k}_1\,$ and $\,\mbold{k}_2\,$ which are emitted from two cells having the same 4--velocity $\,u(x)=u(y)=(1,\mbold{0})\,$. In that case \cite{Neuh86} the $R$--function reduces to \begin{equation} R(k_1,k_2|x,y) = 1 + (\cos\theta)^2 = \sum\limits^2_{(\lambda_1,\lambda_2 = 1)}\Big( \mbold{\epsilon}^{\lambda_1}(k_1) \mbold{\epsilon}^{\lambda_2}(k_2) \Big)^2\;\ , \qquad \label{R:ux=uy} \end{equation} where $\cos\theta=\mbold{k}_1\mbold{k}_2/(|\mbold{k}_1||\mbold{k}_2|)\,$. The physical reason is easily understood for this special case when one considers the last part of equation (\ref{R:ux=uy}). The summation over polarizations (here we choose the Coulomb gauge: $\,\epsilon^{\lambda}(k)=(0,\mbold{\epsilon}^{\lambda}(k))\,$ and $\,\mbold{k}\mbold{\epsilon}^{\lambda}(k)=0\,$) does not just lead to a factor of two as implicitly assumed in \cite{SriKap93a,TimPlumRazWein94,SriKap93,SriKap94}. As illustrated by \figurename~\ref{fig:pol} only one direction of the linear polarization can be chosen equal for both photons, whereas the other polarization directions differ by the angle $\,\theta\,$ between the two momenta. Therefore, the polarization overlap involves $\,(\cos\theta)^2\,$ and is less than 2 for $\,\theta \neq 0\,$ (see (\ref{R:ux=uy})). In a realistic situation the radiation comes of course from many cells with different four--velocities and the full structure of the $Q$--tensor has to be employed. Anyhow, for a non--vanishing angle between the observed photon momenta the correlation is reduced compared to the ``pion inspired'' recipe (\ref{P2:wrong}). \begin{figure} \begin{center} \centerline{\epsfig{file=f_pol.ps,% bbllx=70bp,bblly=580bp,bburx=465bp,bbury=766bp,clip=}} \vskip 3mm \begin{minipage}{\LRcaptionwidth} {\refstepcounter{figure}\label{fig:pol}} \sbox{\tmpbox}{\bf\figurename~\thefigure:\hbox{\ }} \begin{LRdescription}{\usebox{\tmpbox}} \item \footnotesize The linear polarization vectors $\mbold{\epsilon}^{\lambda}(k)$ for two photons with momenta $\mbold{k}_1\,$ and $\mbold{k}_2\,$. \end{LRdescription} \end{minipage} \end{center} \end{figure} For application in dynamical models we recommend to use the following formula: \begin{equation} \label{P2:JmuJnu} P_2(\mbold{k}_1, \mbold{k}_2) = P_1(\mbold{k}_1) P_1(\mbold{k}_2) + \!<\!\!\hat{J}^{\mu_1 \dag}(k_1) \hat{J}^{\mu_2}(k_2)\!\!> <\!\!\hat{J}^{\dag}_{\mu_2}(k_2) \hat{J}_{\mu_1}(k_1)\!\!>\! \end{equation} and calculate the current--current correlator with (\ref{JJ:W}) keeping the explicit form of the $Q$--tensor as given in (\ref{Q:leq}). This is equivalent to (\ref{P2:leq}) but involves only one $\,d^4\!x\,$ integration instead of $\,d^4\!x\,d^4\!y\,$ in (\ref{P2:leq}). The dependence of $\,w(k_1,k_2|x)\,$ on $\,\Delta k = k_1 -k_2\,$ and $\,\bar{k}=(k_1 + k_2)/2\,$ has to be derived from a microscopic model for the photon production. The only general statement which can be made at this point is that due to condition $\,(i)\,$ the current--current correlator decays in the local restframe as function of $\,\Delta k\,$ on the long scale $\,\lambda_s^{-1}\,$. In this paper we do not consider a specific microscopic model for $\,w(k_1,k_2|x)\,$ but rather investigate the role of the tensorial structure of the correlator on photon--photon interferometry. Therefore we write $\,w(k_1,k_2|x) = w(\bar{k},\bar{k}|x) + {\cal O}(\Delta k^{\mu}\Delta k^{\nu})\;$. For simplicity we drop terms of second order in $\,\Delta k\,$. The strength of photon production $\,w(\bar{k},\bar{k}|x)\,$ can to some extend be estimated from the single inclusive cross section \cite{KapLicSeib92}. Based on the discussion above we propose to reexamine the photon intensity interferometry in the spirit of \cite{SriKap93a,TimPlumRazWein94,SriKap93,SriKap94} but using (\ref{P2:JmuJnu}) for the double inclusive cross section. Unfortunately until now the Bose--Einstein correlations of photons have been studied using the formula for double inclusive spectra where the photons are assumed to be scalar massless particles, but normalizing the correlation function to $3/2$ instead of $2$ \cite{SriKap93,SriKap93a,SriKap94}. That formula is just expression (\ref{P2:leq}) with $R(k_1,k_2|x,y) = 2\;$ and reads \begin{equation} \label{P2:wrong} P_2^{wrong}(\mbold{k}_1,\mbold{k}_2) = P_1(\mbold{k}_1) P_1(\mbold{k}_2) + 2 \!\int\!d^4\!x d^4\!y \, \cos(\Delta{}x\Delta{}k) w(\bar{k},\bar{k}|x)w(\bar{k},\bar{k}|y)\;\ . \qquad \end{equation} In order to demonstrate the size of the effect which arises from the vector nature of the photon and the conservation of electric charge we have performed numerical studies of photon intensity interferometry in Bj\o rken hydrodynamics (equation of state $p = \epsilon/3$) using both our formula (\ref{P2:JmuJnu}) (or equivalently (\ref{P2:leq})) and the wrong one (\ref{P2:wrong}). We assume that the photons are produced from the expanding source in local equilibrium and parameterize the photon production rate as $\,w(k_1,k_2|x) \cong w(\bar{k},\bar{k}|x) = N\cdot (T(x))^2 \exp(-(\bar{k}u(x))/T(x))\,$ (for a more precise expression cf. \cite{KapLicSeib92}). The plot of the correlation function defined as $C_2(k_1,k_2) \equiv P_2(k_1,k_2)/P_1(k_1)P_1(k_2) $ shows a significant difference in the results based on the different formulas (see \figurename~\ref{fig:hyd}). The correct expression gives less correlations for non--zero relative rapidities which means that the source size deduced from experimental data is overestimated by the wrong formula (\ref{P2:wrong}). \begin{figure \begin{center} \centerline{\epsfig{file=f_hyd.ps,% bbllx=140bp,bblly=390bp,bburx=500bp,bbury=780bp,clip=}} \vskip 3mm \begin{minipage}{\LRcaptionwidth} {\refstepcounter{figure}\label{fig:hyd}} \sbox{\tmpbox}{\bf\figurename~\thefigure:\hbox{\ }} \begin{LRdescription}{\usebox{\tmpbox}} \item \footnotesize Bose--Einstein correlation plot as a function of rapidity difference $\Delta y$ at fixed transverse momenta $\mbox{\bf k}_1^{\perp}=\mbox{\bf k}_2^{\perp}=100\; \mbox{\rm MeV} $ for Bj\o rken hydrodynamics with $T_i=200\; \mbox{\rm MeV}$, $T_f=140\; \mbox{\rm MeV}$ and initial proper--time $\tau_i=0.3\; fm/c\;$.\\ Solid line corresponds to the photon--photon correlations (\ref{P2:JmuJnu}) (or equivalently (\ref{P2:leq})). \\ Dashed line represents the wrong ``pion--inspired'' expression (\ref{P2:wrong}). \end{LRdescription} \end{minipage} \end{center} \end{figure} In summary we should like to stress that the derivation of a basic equation for the double inclusive spectrum is significantly modified by the fact that photons are massless particles with spin 1 and that they are produced by a conserved electric current. Under the conditions $\,(i)-(iii)\,$ which are fulfilled for high--energy photons in relativistic heavy--ion collisions, the expression (\ref{P2:JmuJnu}) is a suitable starting point for photon intensity interferometry studies. We would like to acknowledge the fruitful discussions with G.~Bertsch and J.~Knoll as well as the comments of D.~Seibert. Especially we are grateful to R.~Weiner and M.~Pl\"umer for critical remarks. \newpage \def\noopsort#1{} \def\SL#1{#1}
\section{Introduction} The decays of $B$ mesons to two charmless hadrons can be described by a $b \rightarrow u$ tree-level spectator diagram (Figure~\ref{fig:feynman}a), or a $b \rightarrow sg$ one-loop ``penguin-diagram'' (Figure~\ref{fig:feynman}b) and to a lesser extent, by the color-suppressed tree (Figure~\ref{fig:feynman}c) or CKM-suppressed $b\rightarrow dg$ penguin diagrams. Although such decays can also include contributions from $b \rightarrow u$ $W$-exchange (Figure~\ref{fig:feynman}d), annihilation (Figure~\ref{fig:feynman}e), or vertical $W$~loop (Figure~\ref{fig:feynman}f) processes, these contributions are expected to be negligible in most cases. Decays such as $B^0 \rightarrow \pi^+\pi^-$ and $B^0 \rightarrow \pi^\pm\rho^\mp$ are expected to be dominated by the $b \rightarrow u$ spectator transition, and measurements of their branching fractions could be used to extract a value for $|V_{ub}|$. The decay mode $B^0 \rightarrow \pi^+\pi^-$ can be used to measure $CP$ violation in the $B$~sector at both asymmetric $B$~factories~\cite{bfact} and hadron colliders\cite{hadcol}. Since the $\pi^+\pi^-$ final state is a $CP$ eigenstate, $CP$ violation can arise from interference between the amplitude for direct decay and the amplitude for the process in which the $B^0$ first mixes into a $\bar{B}^0$ and then decays. Measurement of the time evolution of the rate asymmetry leads to a measurement of $\sin 2\alpha$, where $\alpha$ is one of the angles in the unitarity triangle\cite{unitarity_triangle}. If the $B^0 \rightarrow \pi^+\pi^-$ decay has a non-negligible contribution from the $b \rightarrow dg$ penguin, interference between the spectator and penguin contributions will contaminate the measurement of $CP$ violation via mixing~\cite{pollution}, an effect known as ``penguin pollution.'' If this is the case, the penguin and spectator effects can be disentangled by also measuring the isospin-related decays $B^0\rightarrow \pi^0\pi^0$ and $B^\pm\rightarrow \pi^\pm\pi^0$ \cite{pipicp}. Alternatively, SU(3) symmetry can be used to relate $B^0\rightarrow \pi^+\pi^-$ and $B^0\rightarrow K^+\pi^-$ \cite{silva,oits566}. Penguin and spectator effects may then be disentangled \cite{silva} once the ratio of the two branching fractions and $\sin 2\beta$\cite{unitarity_triangle} are measured. Decays such as $B^0 \rightarrow K^+\pi^-$ and $B^0 \rightarrow K^{*+}\pi^-$ are expected to be dominated by the $b \rightarrow sg$ penguin process, with a small contribution from a Cabibbo-suppressed $b \rightarrow u$ spectator process. Interference between the penguin and spectator amplitudes can give rise to direct $CP$ violation, which will manifest itself as a rate asymmetry for decays of $B^0$ and $\bar{B}^0$~mesons, but the presence of hadronic phases complicates the extraction of the $CP$ violation parameters. There has been discussion in recent literature about extracting the unitarity angles using precise time-integrated measurements of $B$ decay rates. Gronau, Rosner, and London have proposed~\cite{kpicp} using isospin relations and flavor SU(3) symmetry to extract, for example, the unitarity angle $\gamma$ by measuring the rates of $B^+$ decays to $K^0\pi^+$, $K^+\pi^0$, and $\pi^+\pi^0$ and their charge conjugates. More recent publications \cite{deshpande,GRLSU3,EWPrebuttal,EWPmore} have questioned whether electroweak penguin contributions ($b \rightarrow s\gamma$, $b\rightarrow sZ$) are large enough to invalidate isospin relationships and whether SU(3) symmetry-breaking effects can be taken into account. If it is possible to extract unitarity angles from rate measurements alone, the measurements could be made at either symmetric or asymmetric $B$~factories (CESR, KEK, SLAC), but will require excellent particle identification to distinguish between the $K\pi$ and $\pi\pi$ modes. Decays such as $B \rightarrow K\phi$ and $B^+ \rightarrow K^0\pi^+$ cannot occur via a spectator process and are expected to be dominated by the penguin process. Measurement of these decays will give direct information on the strength of the penguin amplitude. Various extensions or alternatives to the Standard Model have been suggested. Such models characteristically involve hypothetical high mass particles, such as fourth generation quarks, leptoquarks, squarks, gluinos, charged Higgs, charginos, right-handed $W$'s, and so on. They have negligible effect on tree diagram dominated $B$ decays, such as those involving $b\rightarrow cW^-$ and $b\rightarrow uW^-$, but can contribute significantly to loop processes like $b\rightarrow sg$ and $b\rightarrow dg$. Since non-standard models can have enhanced $CP$ violating effects relative to predictions based on the standard Kobayashi-Maskawa mechanism \cite{RKM,RNirQuinn}, such effects might turn out to be the key to the solution of the baryogenesis problem, that is, the obvious asymmetry in the abundance of baryons over antibaryons in the universe. Many theorists believe that the KM mechanism for $CP$ violation is not sufficient to generate the observed asymmetry or even to maintain an initial asymmetry through cool-down \cite{RcosmicCP}. Loop processes in $B$ decay may be our most sensitive probe of physics beyond the Standard Model. This paper reports results on the decays $B \rightarrow \pi\pi$, $B \rightarrow K \pi$, $B\rightarrow KK$, $B\rightarrow \pi\rho$, $B\rightarrow K \rho$, $B\rightarrow K^*\pi$, $B \rightarrow K\phi$, $B \rightarrow K^*\phi$, and $B \rightarrow \phi \phi$ \cite{chargeconjugation}. Recent observations of the sum of the two-body charmless hadronic decays $B^0\rightarrow \pi^+\pi^- {\rm and}\ K^+\pi^-$~\cite{kpi} and of the electromagnetic penguin decay $B \rightarrow K^* \gamma$~\cite{kstargamma}, indicate that we have reached the sensitivity required to observe such decays. The size of the data set and efficiency of the CLEO detector allow us to place upper limits on the branching fractions in the range $10^{-4}$ to $10^{-6}$. \section{Data sample and event selection} The data set used in this analysis was collected with the CLEO-II detector~\cite{detector} at the Cornell Electron Storage Ring (CESR). It consists of $2.42~{\rm fb}^{-1}$ taken at the $\Upsilon$(4S) (on-resonance) and $1.17~{\rm fb}^{-1}$ taken at a center of mass energy about 35~MeV below $B\bar{B}$ threshold. The on-resonance sample contains 2.6~million $B\bar{B}$ pairs. The below-threshold sample is used for continuum background estimates. The momenta of charged particles are measured in a tracking system consisting of a 6-layer straw tube chamber, a 10-layer precision drift chamber, and a 51-layer main drift chamber, all operating inside a 1.5 T superconducting solenoid. The main drift chamber also provides a measurement of the specific ionization loss, $dE/dx$, used for particle identification. Photons are detected using 7800 CsI crystals, which are also inside the magnet. Muons are identified using proportional counters placed at various depths in the steel return yoke of the magnet. The excellent efficiency and resolution of the CLEO-II detector for both charged particles and photons are crucial in extracting signals and suppressing both continuum and combinatoric backgrounds. Charged tracks are required to pass track quality cuts based on the average hit residual and the impact parameters in both the $r-\phi$ and $r-z$ planes. We require that charged track momenta be greater than 175~MeV/$c$ to reduce low momentum combinatoric background. Pairs of tracks with vertices displaced from the primary interaction point are taken as $K_S^0$ candidates. The secondary vertex is required to be displaced from the primary interaction point by at least 1~mm for candidates with momenta less than 1~GeV/$c$ and at least 3~mm for candidates with momenta greater than 1~GeV/$c$. We make a momentum-dependent cut on the $\pi^+\pi^-$ invariant mass. Isolated showers with energies greater than $30$~MeV in the central region of the CsI detector, $|\cos\theta| < 0.71$, where $\theta$ is the angle with respect to the beam axis, and greater than $50$~MeV elsewhere, are defined to be photons. Pairs of photons with an invariant mass within two standard deviations of the nominal $\pi^0$ mass \cite{pdb94} are kinematically fitted with the mass constrained to the $\pi^0$ mass. To reduce combinatoric backgrounds we require that the $\pi^0$ momentum be greater than $175$~MeV/$c$, that the lateral shapes of the showers be consistent with those from photons, and that $|\cos\theta^*|<0.97$, where $\theta^*$ is the angle between the direction of flight of the $\pi^0$ and the photons in the $\pi^0$ rest frame. We form $\rho$ candidates from $\pi^+\pi^-$ or $\pi^+\pi^0$ pairs with an invariant mass within $150$~MeV of the nominal $\rho$ masses. $K^*$ candidates are selected from $K^+ \pi^-$, $K^+\pi^0$, $K_S^0\pi^+$ or $K_S^0\pi^0$ pairs \cite{kstars} with an invariant mass within $75$~MeV of the nominal $K^*$ masses. We form $\phi$ candidates from $K^+K^-$ pairs with invariant mass within $6.5$~MeV of the nominal $\phi$ mass. Charged particles are identified as kaons or pions according to $dE/dx$. We first reject electrons based on $dE/dx$ and the ratio of the track momentum to the associated shower energy in the CsI calorimeter. We reject muons by requiring that the tracks not penetrate the steel absorber to a depth of five nuclear interaction lengths. We define $S$ for a particular hadron hypothesis as \begin{equation} S_{\rm hypothesis} = {{dE \over dx}|_{\rm measured} - {dE \over dx}|_{\rm hypothesis} \over \sigma} \end{equation} where $\sigma$ is the expected resolution, which depends primarily on the number of hits used in the $dE/dx$ measurement. We measure the $S$ distribution in data for kaons and pions using $D^0\rightarrow K^-\pi^+$ decays where the $D^0$ flavor is tagged using $D^{*+}\rightarrow D^0\pi^+$ decays. In particular, we are interested in separating pions and kaons with momenta near 2.6 GeV/$c$. The $S_\pi$ distribution for the pion hypothesis is shown in Figure~\ref{fig:dedx} for pions and kaons with momenta between 2.3 and 3.0~GeV/$c$. At these momenta, pions and kaons are separated by $1.8\pm0.1$ in $S_\pi$. \section{Candidate selection} \subsection{Energy Constraint} Since the $B$'s are produced via $e^+e^-\rightarrow \Upsilon (4S)\rightarrow B\bar B$, where the $\Upsilon (4S)$ is at rest in the lab frame, the energy of either of the two $B$'s is given by the beam energy, $E_{\rm b}$. We define $\Delta E = E_1 + E_2 - E_{\rm b}$ where $E_1$ and $E_2$ are the energies of the daughters of the $B$ meson candidate. The $\Delta E$ distribution for signal peaks at $\Delta E =0$, while the background distribution falls linearly in $\Delta E$ over the region of interest. The resolution of $\Delta E$ is mode dependent and in some cases helicity angle dependent (see section III.C) because of the difference in energy resolution between neutral and charged pions. For modes including high momentum neutral pions in the final state, the $\Delta E$ resolution tends to be asymmetric because of energy loss out of the back of the CsI crystals. The $\Delta E$ resolutions for the modes in this paper, obtained from Monte Carlo simulation, are listed in Tables~\ref{tab:deltae} and \ref{tab:mlfits}. We check that the Monte Carlo accurately reproduces the data in two ways. First, the r.m.s.\ $\Delta E$ resolution for $B^0\rightarrow h^+h^-$ (where $h^\pm$ indicates a $\pi^\pm$ or $K^\pm$) is given by $\sigma_{{\Delta E}_{h^+h^-}} = \sqrt{2}\sigma_p$ where $\sigma_p$ is the r.m.s.\ momentum resolution at $p=2.6$\ GeV$/c$. We measure the momentum resolution at $p=5.3$ GeV$/c$ using muon pairs and in the range $p=1.5$--2.5 GeV$/c$ using the modes $B \rightarrow \psi K$, $B\rightarrow D\pi$, and $B \rightarrow D^*\pi$. We find $\sigma_{\Delta E_{h^+h^-}} = 24.7\pm 2.3 ^{+1.4}_{-0.7}$ MeV, where the first error is statistical and the second is systematic. This result is in good agreement with the Monte Carlo prediction. We also test our Monte Carlo simulation in the modes $B^+ \rightarrow \bar D^0\pi^+$ and $B^0\rightarrow D^-\pi^+$ (where $\bar D^0 \rightarrow K^+\pi^-$, $\bar D^0 \rightarrow K_S^0\pi^0$, and $D^-\rightarrow K_S^0\pi^-$) using an analysis similar to our $B \rightarrow K^*\pi$ analysis. Again, $\Delta E$ resolutions for data and Monte Carlo are in good agreement. The energy constraint also helps to distinguish between modes of the same topology. When a real $K$ is reconstructed as a $\pi$, $\Delta E$ will peak below zero by an amount dependent on the particle's momentum. For example, $\Delta E$ for $B\rightarrow K^+\pi^-$, calculated assuming $B\rightarrow\pi^+\pi^-$, has a distribution which is centered at $-42$~MeV, giving a separation of $1.7\sigma$ between $B \rightarrow K^+\pi^-$ and $B \rightarrow \pi^+\pi^-$. \subsection{Beam-Constrained Mass} Since the energy of a $B$ meson is equal to the beam energy, we use $E_{\rm b}$\ instead of the reconstructed energy of the $B$ candidate to calculate the beam-constrained $B$ mass: $M_B = \sqrt{E_{\rm b}^2 - {\bf p}_B^2}$. The beam constraint improves the mass resolution by about an order of magnitude, since $|{\bf p}_B|$ is only $0.3~$GeV/$c$ and the beam energy is known to much higher precision than the measured energy of the $B$ decay products. Mass resolutions range from 2.5 to 3.0 MeV, where the larger resolution corresponds to decay modes with high momentum $\pi^0$'s. Again, we verify the accuracy of our Monte Carlo by studying fully reconstructed $B$ decays. The $M_B$ distribution for continuum background is described by the empirical shape \begin{equation} f(M_B) \propto M_B\sqrt{1-x^2}\exp\left(-\xi(1-x^2)\right) \end{equation} where $x$ is defined as $M_B/E_{\rm b}$ and $\xi$ is a parameter to be fit. As an example, Figure~\ref{fig:argusfcn} shows the fit for $B \rightarrow h^+\pi^0$ background from data taken below $B\bar B$ threshold. \subsection{Helicity Angle} The decays $B \rightarrow \pi\rho$, $B \rightarrow K\rho$, $B \rightarrow K^*\pi$, and $B \rightarrow K\phi$ are of the form pseudoscalar $\rightarrow$ vector + pseudoscalar. Therefore we expect the helicity angle, $\theta_H$, between a resonance daughter direction and the $B$ direction in the resonance rest frame to have a $\cos^2\theta_H$ distribution. For these decays we require $|\cos \theta_H| > 0.5$. \subsection{$D$ Veto} We suppress events from the decay $B^+\rightarrow \bar{D}^0 \pi^+$ (where $\bar{D}^0 \rightarrow K^+\pi^-$ or $\bar{D}^0 \rightarrow K_S^0\pi^0$) or $B^0 \rightarrow D^-\pi^+$ (where $D^-\rightarrow K_S^0\pi^-$) by rejecting any candidate that can be interpreted as $B\rightarrow \bar{D} \pi$, with a $K\pi$ invariant mass within $2\sigma$ of the nominal $D$ mass. We expect less than half an event background per mode from $B\rightarrow \bar D \pi$ events after this veto. The vetoed $D\pi$ signal is used as a cross-check of signal distributions and efficiencies. \section{Background Suppression using Event Shape} The dominant background in all modes is from continuum production, $e^+e^- \rightarrow q\bar{q}\ (q=u,d,s,c)$. After the $D$ veto, background from $b \rightarrow c$ decays is negligible in all modes because final state particles in such decays have maximum momenta lower than what is required for the decays of interest here. We have also studied backgrounds from the rare processes $b \rightarrow s \gamma$ and $b \rightarrow u\ell\nu$ and find these to be negligible as well. Since the $B$ mesons are approximately at rest in the lab, the angles of the decay products of the two $B$ decays are uncorrelated and the event looks spherical. On the other hand, hadrons from continuum $q\bar{q}$ production tend to display a two-jet structure. This event shape distinction is exploited in two ways. First, we calculate the angle, $\theta_T$, between the thrust axis of the $B$ candidate and the thrust axis of all the remaining charged and neutral particles in the event. The distribution of $\cos \theta_T$ is strongly peaked near $\pm1$ for $q\bar q$ events and is nearly flat for $B\bar B$ events. Figure~\ref{fig:costhrcomp} compares the $\cos \theta_T$ distributions for Monte Carlo signal events and background data. We require $|\cos\theta_T| < 0.7$ which removes more than $90\%$ of the continuum background with approximately $65\%$ efficiency for signal events \cite{thrustnotflat}. Second, we characterize the event shape by dividing the space around the candidate thrust axis into nine polar angle intervals of $10^\circ$ each, illustrated in Figure~\ref{fig:vcal}; the $i^{th}$ interval covers angles with respect to the candidate thrust axis from $(i-1)\times 10^{\circ}$\ to $i\times 10^{\circ}$. We fold the event such that the forward and backward intervals are combined. We then define the momentum flow, $x_i$ ($i=1,9$), into the $i^{th}$ interval as the scalar sum of the momenta of all charged tracks and neutral showers pointing in that interval. The $10^\circ$ binning was chosen to enhance the distinction between $B \bar B$ and continuum background events. Angular momentum conservation considerations provide additional distinction between $B\bar B$ and continuum $q\bar q$ events. In $q\bar q$ events, the direction of the candidate thrust axis, $\theta_{q\bar q}$, with respect to the beam axis in the lab frame tends to maintain the $1+\cos^2 \theta_{q\bar q}$ distribution of the primary quarks. The direction of the candidate thrust axis for $B \bar B$ events is random. The candidate $B$ direction, $\theta_B$, with respect to the beam axis exhibits a $\sin^2 \theta_B$ distribution for $B \bar B$ events and is random for $q \bar q$ events. A Fisher discriminant\cite{kendall} is formed from these eleven variables: the nine momentum flow variables, $|\cos \theta_{q\bar q}|$, and $|\cos \theta_B|$. The discriminant, $\cal F$, is the linear combination \begin{equation} {\cal F} = \sum^{11}_{i=1} \alpha_i\ x_i \label{eqn:xfdef} \end{equation} of the input variables, $x_i$, that maximizes the separation between signal and background. The Fisher discriminant parameters, $\alpha_i$, are given by \begin{equation} \alpha_i = \sum_{j=1}^{11} (U_{ij}^b + U_{ij}^{s})^{-1} \times (\mu_j^{b} - \mu_j^{s}). \label{eqn:alphadef} \end{equation} where $U_{ij}^s$\ and $U_{ij}^{b}$\ are the covariance matrices of the input variables for signal and background events, and $\mu_j^s,\ \mu_j^{b}$\ are the mean values of the input variables. We calculate $\alpha_i$ using Monte Carlo samples of signal and background events in the mode $B \rightarrow \pi^+\pi^-$. Figure~\ref{fig:fdstan} shows the $\cal F$ distributions for Monte Carlo signal in the mode $B^0\rightarrow \pi^+\pi^-$, and data signal in the modes $B \rightarrow \bar D\pi$. Figure~\ref{fig:fdstan} also shows the $\cal F$ distributions for Monte Carlo background in the mode $B\rightarrow h^+\pi^-$ and below-threshold background data for modes comprising three charged tracks or two charged tracks and a $\pi^0$. The $\cal F$ distribution for signal is fit by a Gaussian distribution, while the $\cal F$ distribution for background data is best fit by the sum of two Gaussians with the same mean but different variances and normalizations. The separation between signal and background means is approximately 1.3 times the signal width. We find that the Fisher coefficients calculated for $B^0 \rightarrow \pi^+\pi^-$ work equally well for all other decay modes presented in this paper. Figure~\ref{fig:fdmiracle} shows the remarkable consistency of the means and widths of the $\cal F$ distributions for signal and background Monte Carlo for the modes in this study. \section{Analysis} For the decay modes $B \rightarrow \pi\pi$, $B \rightarrow K\pi$, and $B \rightarrow KK$, we extract the signal yield using a maximum likelihood fit. For the other decay modes, we use a simple counting analysis. Both techniques are described below. \subsection {Maximum Likelihood Fit} We perform unbinned maximum likelihood fits using $\Delta E$, $M_B$, $\cal F$, and $dE/dx$ (where appropriate) as input information for each candidate event to determine the signal yields for $B^0\rightarrow \pi^+\pi^-,\ K^+\pi^-,\ K^+K^-,\ \pi^0\pi^0,\ K^0\pi^0$, and $B^+\rightarrow \pi^+\pi^0,\ K^+\pi^0,\ K^0\pi^+$. Five different fits are performed as listed in Table~\ref{tab:mlfits}. For each fit a likelihood function $\cal L$ is defined as: \begin{equation} {\cal L} = \prod_{i=1}^{N} P\left( f_1, ..., f_m; \left(\Delta E, M_B, {\cal F}, dE/dx\right)_i\right) \label{eqn:likefun} \end{equation} where $P\left( f_1, ..., f_m; \left(\Delta E, M_B, {\cal F}, dE/dx\right)_i\right)$ is the probability density function evaluated at the measured point $(\Delta E,\ M_B,\ {\cal F},\ dE/dx)_i$ for a single candidate event, $i$, for some assumption of the values of the yield fractions, $f_j$, that are determined by the fit. $N$ is the total number of events that are fit. The fit includes all the candidate events that pass the selection criteria discussed above as well as $|\cos\theta_T|<0.7$, and $0<{\cal F}<1$. The $\Delta E$ and $M_B$ fit ranges are given in Table~\ref{tab:mlfits}. For the case of $B\rightarrow h^+h^-$, the probability $P_i = P\left( f_1, ..., f_m; \left(\Delta E, M_B, {\cal F}, dE/dx\right)_i\right)$ is then defined by: \begin{eqnarray} P_i &= & f^S_{\pi\pi} P^S_{\pi\pi} + f^S_{K\pi} P^S_{K\pi} + f^S_{KK} P^S_{KK} + (1- f^S_{\pi\pi} - f^S_{K\pi} - f^S_{KK}) P^C \\ P^C &= & f^C_{\pi\pi} P^C_{\pi\pi} + f^C_{K\pi} P^C_{K\pi} + (1- f^C_{\pi\pi} - f^C_{K\pi}) P^C_{KK} \nonumber \end{eqnarray} where, for example, $P^S_{\pi\pi}$ ($P^C_{\pi\pi}$) is the product of the individual probability density functions for $\Delta E$, $M_B$, $\cal F$, and $dE/dx$ for $\pi^+\pi^-$ signal (continuum background). The signal yield in $B^0 \rightarrow \pi^+\pi^-$, for example, is then given by $N_{\pi\pi} \equiv f_{\pi\pi}^S\times N$. The central values of the individual signal yields from the fits are given in Table~\ref{tab:mlresults}. None of the individual modes shows a statistically compelling signal. To illustrate the fits, Figure~\ref{fig:mass_2body} shows $M_B$ projections for events in a signal region defined by $|\Delta E| < 2\sigma_{\Delta E}$ and ${\cal F} < 0.5$ and Figure~\ref{fig:de_2body} shows the $\Delta E$ projections for events within a 2$\sigma$ $M_B$ cut and ${\cal F} < 0.5$. The modes are sorted by $dE/dx$ according to the most likely hypothesis and are shown in the plots with different shadings. Overlaid on these plots are the projections of the fit function integrated over the remaining variables within these cuts. (Note that these curves are not fits to these particular histograms.) Our previous publication \cite{kpi} reported a significant signal in the sum of $B^0 \rightarrow \pi^+\pi^-$ and $B^0 \rightarrow K^+\pi^-$. While our current analysis confirms this result, we now focus on separating the two modes. We separate the systematic errors that affect the total yield from those that affect the separation of the two modes. We do this by repeating the likelihood fit using $N_{\rm sum}\equiv N_{\pi\pi} + N_{K\pi}$, $R \equiv N_{\pi\pi}/N_{\rm sum}$, and fixing $N_{KK} = 0$, its most likely value. We find: \begin{eqnarray*} N_{\rm sum} = && 17.2^{+5.6\ +2.2}_{-4.9\ -2.5} \nonumber \\ R = &&0.54^{+0.19}_{-0.20}\pm 0.05 \end{eqnarray*} where the first error is statistical and the second is systematic (described below). The result of this fit is shown in Figure~\ref{fig:contour}. This figure shows a contour plot (statistical errors only) of $N_{\rm sum}$ {\it vs.}\ $R$ in which the solid curves represent the $n\sigma$ contours ($n=$1--4) corresponding to decreases in the log likelihood by $0.5n^2$. The dashed curve represents the $1.28\sigma$ contour, from which estimates of the 90\% confidence level limits can be obtained. The central value of $N_{\rm sum}$ has a statistical significance of $5.2\sigma$. The significance is reduced to $4.2\sigma$\ if all parameters defining $\cal L$\ are varied coherently so as to minimize $N_{\rm sum}$. Further support for the statistical significance of the result is obtained by using Monte Carlo to draw 10000 sample experiments, each with the same number of events as in the data fit region but no signal events. We then fit each of these sample experiments to determine $N_{\rm sum}$\ in the same way as done for data. We find that none of the 10000 sample experiments leads to $N_{\rm sum}>10$. None of the physical range of $R$\ can be excluded at the $3\sigma$\ level. However the systematic error of $R$ is only 10\% (see below and Table~\ref{tab:mlsyst}). We therefore conclude that our analysis technique has sufficient power to distinguish the $\pi^+\pi^-$ mode from $K^+\pi^-$, but at this time we do not have the statistics to do so. Since none of our fits has a statistically significant signal, we calculate the $90\%$ confidence level upper limit yield from the fit, $N^{90}$, given by \begin{equation} {\int_0^{N^{90}} {\cal L}_{\rm max} (N) dN \over \int_0^{\infty} {\cal L}_{\rm max} (N) dN} = 0.90 \label{eqn:upplim} \end{equation} where ${\cal L}_{\rm max}(N)$ is the maximal $\cal L$\ at fixed $N$\ to conservatively account for possible correlations among the free parameters in the fit. The upper limit yield is then increased by the systematic error determined by varying the parameters defining $\cal L$\ within their systematic uncertainty as discussed below. Table~\ref{tab:mlresults} summarizes upper limits on the yields for the various decay modes. To determine the systematic effects on the yield due to uncertainty of the shapes used in the likelihood fits, we vary the parameters that define the likelihood functions. The variations of the yields are given in Table~\ref{tab:mlsyst}. The largest contribution to the systematic error arises from the variation of the $M_B$ background shape. For this shape, $f(M_B) \propto M_B\sqrt{1-x^2}\exp(-\xi(1-x^2))$ ($x \equiv M_B/E_{\rm b}$), we vary $E_{\rm b}$ by $\pm1$ MeV, consistent with observed variation; we vary $\xi$ by the amount allowed by a fit to background data (below-threshold and on-resonance $\Delta E$ sideband) which pass all other selection criteria. To be conservative, we allow for correlated variations of $E_{\rm b}$ and $\xi$. \subsection {Event-Counting Analyses} In the event-counting analyses we make cuts on $\Delta E$, $M_B$, $\cal F$, and $dE/dx$. The cuts for $\Delta E$ and $M_B$ are mode dependent and are listed in Table~\ref{tab:deltae}. We require ${\cal F} < 0.5$. Tracks are identified as kaons and/or pions if their specific ionization loss, $dE/dx$, is within three standard deviations of the expected value. For certain topologies, candidates can have multiple interpretations under different particle hypotheses. In these cases we use a strict identification scheme where a track is positively identified as a kaon or a pion depending on which $dE/dx$ hypothesis is more likely: we sort the modes with two charged tracks plus a $\pi^0$ ($\pi^+\rho^-$, $\pi^0\rho^0$, $K^+\rho^-$, $K^{*+}\pi^-$, and $K^{*0}\pi^0$) by requiring strict identification for both charged tracks. For modes with three charged tracks ($\pi^+\rho^0$, $K^+\rho^0$, and $K^{*0}\pi^+$) we require strict identification of the two like-sign tracks, while the unlike-sign track \cite{chargestrangeness} is required to be consistent with the pion hypothesis within two standard deviations. We separate modes with one charged track plus two $\pi^0$'s ($\rho^+\pi^0$ and $K^{*+}\pi^0$) by requiring strict identification of the charged track. Figures~\ref{fig:mass_pirho}--\ref{fig:mass_kphi} show $M_B$ distributions for $B \rightarrow \pi\rho$, $B \rightarrow K\rho$, $B\rightarrow K^*\pi$, $B\rightarrow K\phi $, $B\rightarrow K^*\phi$ and $B\rightarrow \phi\phi$ candidates (after making the cuts on $\Delta E$, $\cal F$, and particle identification described above.) The numbers of events in the signal regions are listed in Table~\ref{tab:evcount_results}. In order to estimate the background in our signal box, we look in a large sideband region in the $\Delta E~vs.~M_B$ plane: $5.20 < M_B < 5.27$ GeV and $| \Delta E | < 200$ MeV. The expected background in the signal region is obtained by scaling the number of events seen in the on-resonance and below-threshold sideband regions (weighted appropriately for luminosity). Scale factors are found using a continuum Monte Carlo sample which is about five times the continuum data on-resonance. In many modes, the backgrounds are so low that there are insufficient statistics in the Monte Carlo to adequately determine a scale factor. For these modes, we calculate upper limits assuming all observed events are signal candidates. The estimated background for each mode is also listed in Table~\ref{tab:evcount_results}. Although we find that there are slight excesses above expected background in some modes, no excess is statistically compelling. We therefore calculate upper limits on the numbers of signal events using the procedure outlined in the Review of Particle Properties~\cite{pdb94} for evaluation of upper limits in the presence of background. To account for the uncertainties in the estimated continuum background we reduce the background estimate by its uncertainty prior to calculating the upper limit on the signal yield. \section{Efficiencies} The reconstruction efficiencies were determined using events generated with a GEANT-based Monte Carlo simulation program~\cite{geant}. Systematic uncertainties were determined using data wherever possible. Some of the largest systematic errors come from uncertainties in the efficiency of the $|\cos\theta_T| < 0.7$ cut (6\%), the uncertainty in the $\pi^0$ efficiency (7\% per $\pi^0$), and the uncertainty in the $K_S^0$ efficiency (8\% per $K_S^0$). In higher multiplicity modes, substantial contributions come from the uncertainty in the tracking efficiency (2\% per track). In the $B \rightarrow \pi\rho,\ K\rho,\ K^*\pi$ analyses, the simulation of the efficiency for the particle identification method has a systematic error of 15\%. For the event-counting analyses, the uncertainty in the ${\cal F} < 0.5$ cut is 5\%. The total detection efficiency, ${\cal E}$, is given by ${\cal E} \equiv {\cal E}_r \times {\cal E}_d$, where ${\cal E}_r$ is the reconstruction efficiency and ${\cal E}_d$ is the product of the appropriate daughter branching fractions. The efficiencies, with systematic errors, are listed in Tables~\ref{tab:mlresults} and \ref{tab:eff}. \section{Upper Limit Branching Fractions} Upper limits on the branching fractions are given by $N_{\rm UL}/({\cal E} N_B)$ where $N_{\rm UL}$ is the upper limit on the signal yield, ${\cal E}$ is the total detection efficiency, and $N_B$ is the number of $B^0$'s or $B^+$'s produced, 2.6 million, assuming equal production of charged and neutral $B$ mesons. To conservatively account for the systematic uncertainty in our efficiency, we reduce the efficiency by one standard deviation. The upper limits on the branching fractions appear in Tables \ref{tab:mlresults} and \ref{tab:evcount_results}. \section{Summary and Conclusions} We have searched for rare hadronic $B$ decays in many modes and find a signal only in the sum of $\pi^+\pi^-$ and $K^+\pi^-$. The combined branching fraction, ${\cal B}(\pi^+\pi^- + K^+\pi^-) = (1.8^{+0.6+0.2}_{-0.5-0.3}\pm0.2) \times 10^{-5}$, is consistent with our previously published result. We have presented new upper limits on the branching fractions for a variety of charmless hadronic decays of $B$~mesons in the range $10^{-4}$ to $10^{-6}$. These results are significant improvements over those previously published. Our sensitivity is at the level of Standard Model predictions for the modes $\pi^+\pi^-,\ K^+\pi^-,\ \pi^+\pi^0,\ K^+\pi^0,\ \pi^{\pm}\rho^{\mp},\ K^+\phi$,\ and $K^{*0}\phi$. \acknowledgments We gratefully acknowledge the effort of the CESR staff in providing us with excellent luminosity and running conditions. J.P.A., J.R.P., and I.P.J.S. thank the NYI program of the NSF, G.E. thanks the Heisenberg Foundation, K.K.G., M.S., H.N.N., T.S., and H.Y. thank the OJI program of DOE, J.R.P thanks the A.P. Sloan Foundation, and A.W. thanks the Alexander von Humboldt Stiftung for support. This work was supported by the National Science Foundation, the U.S. Department of Energy, and the Natural Sciences and Engineering Research Council of Canada.
\section{INTRODUCTION} In recent years the Compton Gamma Ray Observatory (CGRO) and other instruments have provided major new discoveries and detailed observations of isolated $\gamma$-ray pulsars, including the Crab (Nolan et al. 1993), Vela (Kanbach et al. 1994), Geminga (Halpern \& Holt 1992, Bertsch et al. 1992, Mayer-Hasselwander et al. 1994), PSR B1509-58 (Wilson et al. 1992), PSR B1706-44 (Thompson et al. 1992), PSR B1055-52 (Fierro et al. 1993), and most recently PSR B1951+32 (Ramanamurthy et al. 1995). Models of these objects must now account for a variety of detailed features in the emission, especially from the most intense sources (Crab, Vela, Geminga). Current models have in fact already encountered problems in explaining how these sources can show both remarkable similarities and puzzling variations in their light curves and phase-resolved energy spectra. These difficulties are even more severe if models of the $\gamma$- ray emission must also be consistent with the radiation observed at radio, optical, and X-ray wavelengths. As the observational statistics for the weaker sources improve, these theoretical challenges may become even more formidable. At present two general types of $\gamma$- ray pulsar models are popular in the literature. The Polar Cap (PC) model, first proposed by Sturrock (1971) and later investigated by numerous authors (see for example Ruderman and Sutherland 1975, Harding 1981, Daugherty and Harding 1982, Arons 1983) assumes that the emission is produced by electrons accelerated to high energies just above the surface of a magnetized rotating neutron star (NS), in the vicinity of the magnetic poles. In contrast, the Outer Gap model (Cheng, Ho, and Ruderman 1986a, 1986b) places the acceleration regions much higher in the NS magnetosphere, in vacuum gaps formed within a charge-separated plasma. In a previous paper (Daugherty and Harding 1994, hereafter DH94) we proposed a version of the PC model based on the following principal assumptions: (a) The gamma emission is initiated by the acceleration of electrons from the NS surface, just above the magnetic PC regions which enclose the {\it open\/} magnetic field lines extending to the velocity-of-light cylinder (LC). (b) The emission originates as curvature radiation (CR) produced by the electrons as they follow the curvature of the open magnetic field lines. (c) The processes of direct $1-\gamma$ pair conversion (see for example Erber 1966) by the NS magnetic field and synchrotron radiation (SR) by the emitted pairs produce photon-pair cascades, from which the observed $\gamma$ radiation emerges. (d) The rotational and magnetic axes of the radiating NS are nearly aligned, so that the inclination $\alpha$ is small enough to be comparable with the PC half-angle $\theta_{pc}$. More precisely, the model requires that $\alpha \sim \theta_b$ where $\theta_b$ is the half-angle of the $\gamma$-beam emerging from the PC. Assumptions (a)-(c) comprise essentially the original postulates of the PC model (Sturrock 1971). They describe the overall physics of the cascade process and in combination they determine the form of the production spectra for the gamma rays and the pairs. The final assumption (d) primarily affects the viewing geometry. It implies that randomly oriented observers should see emission from at most one PC. However, since CR-induced cascades are intrinsically hollow-cone sources which produce their most intense emission near the PC rim, observers viewing a single PC may detect light curves with either single or double peaks (DH94). Sterner and Dermer (1994) independently noted a similar effect in a model of PC cascades initiated by Comptonization rather than CR. The assumption that $\alpha \sim \theta_b$ allows the phase separation between double peaks to become large enough to match the observed values ($\sim 0.4$ for the Crab and Vela, $\sim 0.5$ for Geminga). In the present work we refine assumption (d) by requiring {\it only\/} that $\alpha \sim \theta_b$, not that $\alpha$ itself be necessarily small. Hence in place of the Nearly Aligned Rotator (NAR) model described in DH94, we consider here a more general Single Polar Cap (SPC) model. In addition, we introduce a further assumption which allows $\theta_b$ (and $\alpha$) to have significantly larger values than $\theta_{pc}$ itself: (e) the acceleration of the electrons occurs over an extended distance above the PC surface, so that they reach their peak energies at heights of a few NS radii. Above these heights, the acceleration is cut off by an overlying force-free plasma. In DH94 we neglected the height of the acceleration region and simply supplied the electrons with an injection energy at the NS surface, then traced their CR energy losses as they escaped outward along field lines for which ${\bf E \cdot B} \sim 0$. We have since noted that the assumption of an extended acceleration region provides a solution to a serious difficulty with our previous model, namely the ``observability'' problem. This refers to the fact that if conventional estimates of PC dimensions are accurate, $\gamma$-beams emitted by energetic electrons just above the NS surface would be so small that there would be a low probability ($\lesssim 10^{-2}$) that they could be detected by randomly oriented observers. In our previous work we noted that the usual estimate for the PC radius $R_{pc}$ may in fact be too small, although moderate increases in $R_{pc}$ cannot by themselves resolve the observability problem. However, the outward flaring of the magnetic field lines implies that the half-angle $\theta_{b}$ of the (hollow) cascade $\gamma$-beams increases rapidly with height above the NS surface. Thus the effect of extending the acceleration zone up to heights of a few NS radii, especially if combined with moderately increased ($\lesssim 2$) PC dimensions, can produce rotating beams whose edges sweep over a much larger solid angle. In DH94 we also noted that in order to produce double peaks as narrow as those observed from the Crab, Vela, and Geminga, we had to assume that the surface density of the electrons drawn from the NS surface is concentrated near the PC rim. In the present work we suggest a physical basis for this empirical observation, namely the acceleration of secondary cascade electrons created near the rim. More precisely, the excess rim density may be supplied by a multistep process initiated by the reversed acceleration of secondary positrons created just below the acceleration cutoff height. These particles can produce downward-oriented cascades, creating new pairs near the NS surface. A fraction of the electrons from these pairs may then be accelerated upward along with the true primary electrons, adding to the net outward flow. We argue that this sort of cascade feedback process should occur preferentially near the PC rim, where the open magnetic field lines have their maximum curvature. \newpage \section{EXTENDED PAIR CASCADES} In our treatment of the NAR Model in DH94, we assumed that the acceleration of PC electrons starts at the NS surface and is cut off sharply at a height $h \ll R_{pc}$ by an overlying force-free pair plasma. This assumption was made partly for simplicity, and also because there is still no firmly established, self-consistent electrodynamical model for magnetospheric acceleration, either near the NS surface or elsewhere. However, we note that a significant problem with these models may be resolved if the accelerating potential $\Phi(h)$ extends upward to heights $h \sim 2 \-- 3$ NS radii or higher. We will first discuss our motivation for exploring extended acceleration regions, then describe our model results based on specific empirical choices for $\Phi(h)$. For simplicity we retain our NAR model assumption that each PC is almost circular with radius $R_{pc} = R_{ns} \theta_{pc}$, where $R_{ns}$ is the NS radius. While the more general SPC model allows larger values of $\alpha$ and hence noncircular PCs, this approximation should be still adequate for our present treatment. For purely dipolar fields, the conventional estimate for the half-angle $\theta_{pc}$ is just \begin{equation} \label{tpc} \sin\theta_{pc} = \left({R_{ns} \over R_{lc}}\right)^{1/2} = \left({R_{ns} \Omega\over c}\right)^{1/2} \end{equation} where $R_{lc} = c/\Omega$ denotes the distance to the velocity-of-light cylinder and $\Omega$ is the NS angular rotation frequency. Eq. (\ref{tpc}) assumes that a dipole field line, emerging from a point near the PC rim, should close just inside the light cylinder. However, as we noted in DH94 this estimate ignores all plasma effects and thus should be regarded only as a lower limit on $\theta_{pc}$. For example, Michel (1982, 1991) has found that the presence of a force-free, rigidly corotating plasma (even without inertial effects or outward current flow) causes a distortion of the field lines which increases the PC radius by a factor $\sim 1.3$. Hence we argue that a more realistic model could be expected to increase $\theta_{pc}$ by a factor $\sim 2$ over Eq. (\ref {tpc}). If we make the usual assumption that the magnetic field is purely dipolar, the equation describing a given field line emerging from the PC is just \begin{equation} \label{dpl} r = k \sin^{2}\theta \end{equation} where k is constant. At a given point on the field line, the angle $\psi$ of the local tangent (measured from the magnetic axis) is given by \begin{equation} \label{tnl} \tan\psi = {{3 \sin\theta \cos\theta} \over {3 {\cos^{2}\theta- 1}}} \end{equation} If the gamma beam size is approximately determined by the locus of tangents to the outermost open field lines, for $\theta_{pc} \ll 1$ a cascade gamma beam originating from the NS surface would have a half angle $\theta_{b} \sim \tan\psi \sim {3 \over 2} \theta_{pc}$. In general we can use Eqs. (\ref{dpl}) and (\ref{tnl}) to estimate the increase in beam width $\theta_{b} \sim \psi$ with height, for a given PC radius. Figure 1 illustrates this height dependence by plotting the tangent angle $\psi$ vs. radial distance along the field lines, for the case of the Vela pulsar ($P = 0.89$ ms). The curves labeled 1, 2,... denote field lines emerging from the NS surface at the corresponding multiples of $\theta_{pc}$ as given by Eq. (\ref{tpc}). It is evident that if the cascade gamma emission extends upward to heights exceeding $\sim 3$ NS radii, $\theta_{b} \sim \psi$ can become significantly larger than ${3 \over 2} \theta_{pc}$. This effect is even more pronounced if $\theta_{pc}$ is taken to be $\sim 2$ or more times the standard estimate (\ref{tpc}). \section{ACCELERATION AND ENERGY LOSSES ABOVE POLAR CAPS} We have shown that from the standpoint of viewing geometry, extended PC cascades may provide a viable solution to the observability problem. The obvious next step is to examine the possiblity that the acceleration of electrons from the PC surface might be sustained up to heights of several NS radii. This question also requires us to consider in detail the energy loss mechanisms which may affect the net acceleration. Due to the intense ($\sim 10^{12} G$) NS magnetic fields, electrons accelerated from the PC surface are constrained by rapid SR losses to follow the field lines. Hence they obey a one-dimensional equation of motion, which may be expressed as an energy-balance equation: \begin{equation} \label{erg} {d\gamma \over ds} = (\beta c)^{-1} \left[ \left({d\gamma \over dt}\right)_{acc} - \left({d\gamma \over dt}\right)_{cr} - \left({d\gamma \over dt}\right)_{cs} - \left({d\gamma \over dt}\right)_{other} \right] \end{equation} Here $\gamma$ denotes the electron Lorentz factor, $\beta = v/c$, and $s$ is the distance traversed along the field line. The subscripts labeling the component energy gain and loss rates are defined as follows. The subscript {\it acc\/} denotes the energy gain due to electrostatic acceleration in regions where ${\bf E \cdot B}$ is nonzero. We assume this term is proportional to ${\bf E_{\parallel}}$, the component of ${\bf E}$ parallel to ${\bf B}$, at each point along the particle trajectory (magnetic field line). Unfortunately, current models of pulsar magnetospheres do not agree on the behavior of ${\bf E_{\parallel}(r)}$ near the PC surface. Hence the energy-gain term in Eq. (\ref{erg}) must be regarded as unknown. However, we can at least assume various simple models for the accelerating potential (e.g. Ruderman and Sutherland 1975, Arons 1983) in our simulations and compare the results for each model with observations. In Sections 6 and 7 we show that we have been able to find self-consistent models of extended cascades which yield light curves and spectra similar to the observed values. We have also identified significant constraints on the accelerating field which are critical to the viability of these models. In contrast to the gain-rate term, the principal loss-rate terms in Eq. (\ref{erg})are reasonably well understood. The subscripts {\it cr\/}, {\it cs\/}, and {\it other\/} denote energy losses due to CR, Compton upscattering (Dermer 1990, Chang 1995), and other scattering processes respectively. One example of the latter is triplet pair production (Mastichiadis et al. 1986, Mastichiadis 1991, Dermer and Schlickeiser 1991). Sturner (1995) has recently provided a systematic treatment of PC electron acceleration which considers these energy-loss processes in detail. We have used his expressions for the CS loss terms in our simulations, although his treatment involves a number of simplifying approximations. The CR loss rate has the simple form (see for example Jackson 1975) \begin{equation} \label{crl} \left({d\gamma \over dt}\right)_{cr} = {2 \over 3} {e^2 \over {m c}} {\gamma^4 \over {\rho_{c}}^2} \end{equation} where $\rho_{c}$ is the local radius of curvature of the magnetic field line. For a purely dipolar field, the exact expression for $\rho_{c}$ is just \begin{equation} \label{crd} \rho_{c} = {{k (\sin^4 \theta + \sin^2 2 \theta)} \over {\sin^4 \theta + 2 \sin^2 2 \theta - 2 \sin^2 \theta \cos 2 \theta}} \end{equation} Since Eq. (\ref{crd}) yields values $~\sim 10^7$ cm for standard PC model parameters, the CR loss rate only becomes significant for $\gamma \gtrsim 10^6$. At higher energies it is by far the dominant loss mechanism. The CS loss rate results from upscattering of ambient photons by the accelerated electron beam. In our model the photon background consists of thermal emission from the NS surface, and hence the CS loss rate should only be significant only at heights $h \lesssim R_{ns}$ above the surface. Pulsed X-ray observations of Geminga (Halpern and Ruderman 1993) and PSR B1055-52 (\"Ogelman and Finley 1993) suggest that for at least some sources the thermal background may include multiple components at distinct temperatures (e.g., emission from both the cooling NS surface and hotter regions in the vicinity of the PCs). The CS loss rate is found from the general expression (Dermer 1990, Sturner 1995) \begin{equation} \label {cs1} \left({d\gamma \over dt}\right)_{cs} = c \int d\epsilon \int d\Omega n_{ph}(\epsilon, \Omega) (1 - \beta \cos \Psi) \int d\epsilon'_s \int d\Omega'_s {d\sigma' \over {d\epsilon'_s d\Omega'_s}} (\epsilon_s - \epsilon) \end{equation} where $\epsilon = \hbar \omega / {m_e c^2}$ is the incident photon energy in units of the electron rest energy, $n_{ph}(\epsilon, \Omega)$ is the number density of incident background photons within energy and solid-angle increments $d\epsilon$ and $d\Omega$, and $\Psi$ denotes the angle between these photons and the local electron beam direction. The quantity $d\sigma' / d\epsilon'_s d\Omega'_s$ is the magnetic Compton scattering cross section in the local electron rest frame (ERF), where the primes denote quantities evaluated in the ERF and the subscript $s$ labels scattered photon quantities. In the strong magnetic field the CS cross section includes both nonresonant and resonant components (Herold 1979, Daugherty and Harding 1986, Bussard et al. 1986). Dermer (1990) has derived a nonrelativistic approximation for the loss rate based on the magnetic Thomson cross section in the ERF (Herold 1979), resolving the total loss rate into component terms which he labels `angular`, `nonresonant`, and `resonant`. Sturner (1995) has applied further simplifying assumptions to these terms in order to derive convenient expressions for the CS loss rate. His results are summarized in his equations (4)-(9), which we have incorporated into our acceleration tracing algorithm. Sturner (1995) notes that for $\gamma \gtrsim 10^3$, the incident thermal photon energies above the cyclotron resonance may become relativistic $(\epsilon' \gtrsim 1)$. In this case he replaces the nonresonant component of the cross section by a relativistic (but nonmagnetic) Klein-Nishina expression given by his equations (10)-(14). In this work we have included these expressions, although we note that a more accurate treatment will require the use of the magnetized (resonant) Klein-Nishina cross section (Daugherty and Harding 1986, Bussard et al. 1986). The only loss term which Sturner (1995) includes under the `other` label in Eq. (\ref {erg}) arises from electron-photon scattering events in which the scattered photon is replaced by an emergent $e^+/e^-$ pair. Using cross sections found by Mastichiadis et al. (1986) and Mastichiadis (1991) for the nonmagnetic form of this process, Sturner (1995) applies a monoenergetic photon approximation to derive a loss rate given by his equation (16). For our model parameters this term is never dominant, but for generality we have also included it in our simulation. As in the case of his Klein-Nishina CS loss rate, however, we note that in future work the magnetic form of this process should be investigated since it may also exhibit resonant behavior which may increase its signifance. \newpage \section{SIMULATIONS OF EXTENDED PC CASCADES} The basic features of our cascade simulation code are described in DH94. The version used in this work includes several major improvements. These include revised adaptive algorithms for tracing photon propagation, which allow more accurate localization of near-threshold pair conversion events. We have also improved the tracing of synchrotron/cyclotron emission, which now more accurately simulates both recoil and angular distribution effects in the cyclotron regime ($\gamma \gtrsim 1$). However, the most significant improvement for this work is the algorithm for tracing electron acceleration through extended regions above the PC. In our current version, each primary electron emerges from the surface with an initial Lorentz factor $\gamma_0 \gtrsim 1$. Assuming specific parameters for both the energy gain and loss mechanisms as described in Section 3 above, the calculation then traces the net acceleration of the electron as it escapes outward along the local magnetic field line. For this purpose we have developed an adaptive numerical technique to integrate Eq. (\ref {erg}) which accomodates a wide range of energies and distance scales. To estimate the significance of Compton losses due to thermal photons from the NS surface we have used a model similar to that employed by Sturner (1995), in which the PC has a uniform surface temperature $T_6$ (in units of $10^6 K$) within a circle of radius $R_{tpc}$ centered on each magnetic pole. This region is defined as the {\it thermal\/} PC. Note that $R_{tpc}$ may differ from the PC radius as defined by the locus of the outermost open field lines. In fact we treat both $T_6$ and $R_{tpc}$ as parameters in the model. At present we ignore any softer emission which may be emitted from the overall surface. Figure 2 shows sample acceleration profiles $\gamma(h)$, where $h$ is the height above the NS surface in stellar radius units. Curve (a) shows a case in which the accelerating field is assumed to be constant, namely $(d\gamma / ds)_{acc} = 5$ ${\rm cm}^{-1}$, from the surface up to a sharp cutoff at height $h_c = 3$. Curves (b) and (c) both assume that the gain rate is a linearly increasing function $(d\gamma / ds)_{acc} = 5h$, over this same region. They differ only in the assumed values for the Comptonization parameters, namely the thermal PC temperature $T_6$ and radius $R_{tpc}$ (measured in NS radius units). Curve (b) assumes $T_6 = 1$ and $R_{tpc} = 0.1$, corresponding to a cool, small thermal PC. The opposite case of a hot, large PC, is shown by curve (c) which assumes $T_6 = 2$ and $R_{tpc} = 0.5$. We note that the constant-acceleration curve (a) is not affected by these variations of the Comptonization parameters, since in this case the gain rate greatly exceeds the loss terms in Eq. (\ref{erg}). We also observe that even the linear-acceleration profiles are sensitive to the Comptonization parameters only for heights $h \ll 1$, and they have little effect on the peak energies reached at the cutoff height $h_c$. As $\gamma$ exceeds values $\gtrsim 10^6$ the primary CR emission reaches gamma ray energies, resulting in photon-pair cascades. The calculation, as described in DH94, recursively traces the full cascade development and accumulates 3D tables of emergent $\gamma$-ray counts vs. energy and solid angle, from which we derive spectra and light curves of the emission as seen from arbitrary viewing directions. In this work we have accumulated photon counts from ensembles of primary electrons distributed in concentric rings over the PC surface. We have assumed that the primary beam current is axisymmetric with respect to the magnetic axis, hence the electrons in each ring are spaced uniformly in azimuth. However, our analysis facility allows us to assign arbitrary weights to the $\gamma$ counts from each ring. This technique allows us to vary the assumed radial dependence of the primary electron current density without requiring new runs of the simulation. Finally we should point out that our current simulation is based strictly on a CR-initiated cascade, i.e. it considers Comptonization as an energy loss mechanism acting on the primary electrons but it does not yet include the upscattered photons as a source of high-energy $\gamma$-rays which may themselves initiate cascades. This is in obvious contrast to the cascade model proposed by Sturner and Dermer (1994), in which Comptonization provides {\it all \/} the high-energy input photons. Under our model assumptions the primaries reach much higher peak energies ($\gamma \gtrsim 10^6$) than the values they assume ($\gamma \sim 10^5$), so that in our case CR should initiate the bulk of the cascade emission. However, we recognize that Comptonization may add a measurable contribution to the emergent $\gamma$-emission and in a separate work we will extend the cascade simulation to trace the CS upscattered photons as well. At the same time, we note that the CS contribution may be expected to produce a narrower $\gamma$-beam than the extended CR component we consider here, since it should originate closer to the PC surface. Thus it is possible that PC cascades initiated by CR and CS photons may be distinguishable both spatially and energetically. \newpage \section{ELECTRON CURRENTS NEAR THE PC RIM} In DH94 we showed that single magnetic poles can exhibit doubly peaked light curves with phase separations $\delta\phi \lesssim 0.5$ if $\alpha \sim \theta_b$ and the observer angle $\zeta \sim \alpha$. However, in order to reproduce the small duty cycles of the double peaks seen in the Crab, Vela, and Geminga, we had to impose an additional {\it ad hoc\/} assumption that the primary electron density is strongly concentrated near the PC rim. We also noted that there are two possibilities for obtaining doubly peaked profiles with $\delta\phi < 0.5$, in which the designations of {\it leading\/} and {\it trailing\/} peaks are reversed. In DH94 we considered in detail the case in which the first peak corresponds to the phase at which the observer viewpoint emerges from the interior of the (hollow) $\gamma$-beam, while the second peak marks the point of reentry. This case, which we denote as the Exterior Scenario (ES), can produce $\delta\phi < 0.5$ if the rotational axis is contained within the $\gamma$-beam ($\alpha < \theta_{b}$). By combining the ES with the assumption that the primary current is concentrated near the PC rim, we could account for both the short duty cycles and the lack of emission outside the peaks (since this would be the phase interval during which the observer viewpoint penetrates the interior of the hollow beam). In this scenario we associated the finite emission observed between the peaks with residual, higher-altitude cascades, which would produce emission with larger beam widths. In work following DH94 we have compared our model predictions in detail with CGRO observations of phase-resolved spectra for the Vela pulsar (Kanbach et al. 1994). We have concluded that the ES does not provide uniformly consistent fits to the spectra, especially for the phase intervals between the main peaks. In the ES model the high-altitude cascades which produce the interpeak emission do tend to produce harder spectra below their characteristic high-energy turnovers, since a smaller fraction of the hard CR emission is converted to softer cascade photons. By itself this trend is at least qualitatively consistent with the observations. However, the peak CR energy ($\propto \gamma^3$) also decreases rapidly as the primaries lose energy above the acceleration zone, with the result that the turnovers in the interpeak cascade spectra drop to lower energies compared to the peak spectra. In this respect the model prediction is opposite to the observed trend. This problem with the ES has led us to reexamine the alternative labeling of the leading and trailing peaks, in which the PC interior is identified as the source of the interpeak emission. We refer to this case as the Interior Scenario (IS). In order to produce finite interpeak emission in this case, we must abandon the phenomenological DH94 model of a pure rim distribution for the primary electrons. However, if we replace the pure rim model with a two-component model which includes a uniform interior current, it turns out that the IS allows a more consistent overall agreement with the observations than the ES. Moreover, in this scenario we can suggest a tentative physical interpretation for a two-component primary current. In particular, the uniform component is a simple approximation of a Goldreich-Julian (GJ) current $I_{GJ} = \pi {R_{pc}}^2 c \rho_{0}$ (Goldreich and Julian 1969), where \begin{equation} \label{gjc} \rho_{0} \sim {{-\bf{\Omega \cdot B}} \over {2 \pi c}} \end{equation} which should be valid if $\theta_{pc} \ll 1$. We propose that this component includes all the true {\it primary\/} electrons drawn from the NS surface. In this view the extra rim component consists of secondary electrons from pairs preferentially created near the PC rim, where the increasing field-line curvature produces more rapid $\gamma$-pair conversions. If any secondary pairs contribute to the PC current of high-energy particles which initiates cascades, the pairs must themselves be accelerated to energies comparable with the peak primary energies. This in turn would require at least some pairs to be created well below the acceleration cutoff height. If (as we assume here) the primaries are negative electrons ($e^{-}$), each $e^{-}$ secondary would then move {\it upward\/} and thus add to the GJ primary current, while the $e^{+}$ would be accelerated {\it downward\/} along the local field line toward the surface. In fact the model $\gamma$-ray light curves we present in Section 6 show that if just a small fraction ($\sim 10^{-2}$ or less) of the cascade pairs created near the rim can be boosted to $\gamma \gtrsim 10^6$, a two-component current model shows good agreement with observations. In spite of these results, we must first consider a fundamental theoretical objection to the acceleration of secondary pairs. The problem is that the onset of cascade pair production is expected to produce a sharp cutoff in the acceleration of the primaries at a height $h_c$, which marks the boundary of the overlying pair plasma (e.g. Ruderman and Sutherland 1975, Arons 1983). Our own simulation results confirm that the quenching of $E_{\parallel}$ above $h_c$ should be an abrupt process, since the density of created pairs is found to rise sharply with height. This is demonstrated in Figure 3, which plots typical growth curves of the multiplicity $M = (N_s^+ + N_s^-)/N_p$ where $N_p$ and $N_s$ denote the numbers of primaries and secondaries respectively. Thus even if pairs created at the lowest heights can be accelerated by a decreasing $E_{\parallel}$ within a finite transition zone, the growth curves indicate that this zone is too short for any $e^{-}$ secondaries to reach energies comparable to those of the primaries. This appears to eliminate the most obvious model for enhancing the PC current near the rim, in which the negative pair members are accelerated outward with the primaries. However, the positron ($e^{+}$) component in such a transition zone must also be subject to acceleration. The key point here is that these particles may be drawn {\it downward\/} from the transition zone back into the acceleration zone, following the local field lines back toward the NS surface. In fact they should traverse a distance comparable to the full extent of the acceleration zone, and thus reach energies sufficient to create (tertiary) pairs by a variety of possible mechanisms (e.g. $\gamma-B$ pair production, triplet pair production). The result would be the creation of pairs deep within the acceleration zone, whose $e^{-}$ members could be accelerated outward with the primaries to reach similar peak energies. This sort of cascade feedback process should be most likely to occur above those regions of the PC where the original upward-directed cascades initiated by the primaries commence at the lowest heights. Unless the electrostatic acceleration varies greatly over the PC interior regions, the increasing curvature of the field lines from the pole to the rim implies that the primary cascades develop first near the rim (cf. Figure 3). Hence we argue that reverse $e^+$-acceleration and downward-oriented cascades occur preferentially around the rim. As a first test of this hypothesis we have generalized our acceleration tracing algorithm to follow secondary positrons downward from creation points just below the cutoff height, back toward the NS surface. The results confirm that these particles can be boosted to $\gamma \lesssim 10^7$ at heights $h \gtrsim R_{ns}$ above the surface, allowing their CR spectra to reach pair-conversion energies and initiate downward-oriented cascades. In a separate work we will refine our complete simulation code to investigate the development of these cascades in detail. We anticipate that their presence may impact our model in several respects, since in addition to providing a new source of electrons these cascades can influence the behavior of the acceleration process just above the surface. In particular, if the cascades create a sufficiently dense layer of pair plasma overlying the surface they can retard acceleration below the effective height of this layer. In addition, it is possible that energetic downstreaming cascade photons can impose severe Comptonization losses on upward-directed electrons. As described in the following sections, in this work we will allow for these possibilities by considering simple models in which the acceleration may effectively commence at finite heights above the NS surface. \section{GAMMA-RAY LIGHT CURVES} The 3D photon count tables accumulated by the simulation may be summed over energy bins to produce 2D sky maps of the $\gamma$-emission between arbitrary energy limits. An example is shown in Figure 4, which plots a grayscale contour map of emission over 100 MeV. Any horizontal line drawn across this plot corresponds to a specific value of the polar angle $\zeta$ for a given viewing direction, and the counts distributed along this line define the $\gamma$-ray light curve as seen from this viewpoint. Following the arguments in Section 5 we present sample results for the Vela pulsar using a simple two-component primary current model, which we obtain by superimposing simulation datasets for concentric rings of primaries as discussed in Section 4. In each case we have included a total of 10 rings spaced at equal radial increments to cover the PC interior. Since each ring contains 180 particles with a uniform 2-degree azimuthal spacing, the inner rings are weighted $\propto r^{-1}$ to approximate a uniform interior density. To simulate test cases with a moderate rim component, we have weighted the outermost ring by arbitrary factors in the range 3 to 5. Physically this corresponds to the acceleration of a few secondary electrons for each primary electron on this ring, which is a small fraction of the $10^{2}-10^{3}$ cascade pairs created per primary near the rim. All the datasets we have accumulated to date assume the following general form for the accelerating field, namely \begin{equation} \label{acc1} E_{\parallel}(h) = {{m c^2} \over e} \left({d\gamma \over ds}\right)_{acc} = {{m c^2} \over e} [a_0 + a_1 (h - h_0)] \Theta(h - h_0) \Theta(h_c - h) \end{equation} We choose units for Eq. (\ref{acc1}) such that the path length $s$ is measured in cm, while the height $h = (R - R_{ns})/R_{ns}$ is in NS radius units from the PC surface, $\Theta (x)$ is the unit step function ($0$ for $x < 0$, $1$ for $x > 0$), and the constants $a_0$, $a_1$, and $h_0$ are taken as free parameters in our model. Their values effectively determine the height at which cascades commence above the PC rim, which Arons (1983) denotes as the ``pair formation front''. In the following we take the height at which the cascade multiplicity exceeds unity (cf. Figure 3) as a reasonable measure of the acceleration cutoff height $h_c$. Thus $h_c$ is a function of $(a_0, a_1, h_0)$ but is {\it not\/} itself a free parameter. In practice we determine $h_c$ from trial simulations before generating complete datasets. The quantity $h_0 \ge 0$ in Eq. (\ref{acc1}) denotes the height at which acceleration commences. We introduce $h_0$ to allow for the possibility that downward-oriented cascades may prevent or impede acceleration just above the NS surface. As noted in Section 5, this can occur either if the cascades create a sufficiently dense layer of pair plasma overlying the surface, or if downstreaming cascade photons impose strong Comptonization losses on upward-moving electrons. In a separate study of downward-oriented cascades we will investigate both of these effects in order to put physical constraints on the choice of $h_0$, but here we simply explore the effects of varying $h_0$ in sample models. If we let $a_1 = 0$ in Eq. (\ref{acc1}) we obtain a constant-field approximation, resembling vacuum gap acceleration models of the type proposed by Ruderman and Sutherland (1975). If instead we set $a_0 = 0$, we have a crude approximation for the potential suggested by Arons (1983) in his slot-gap model. We have generated datasets for the Vela pulsar using both of these limiting forms. In each case we have empirically chosen combinations of the parameters $(a_0,a_1,h_0)$ such that the primary electrons reach their peak energies ($\gamma \gtrsim 10^6$) rapidly enough to initiate cascades. We note that in this work we have assumed no dependence of either $a_0$, $a_1$, or $h_0$ on the magnetic polar angle $\theta$. We have used a further simplifying assumption here, namely that the cutoff height $h_c$ has the same value over the PC interior as determined by the onset of cascades near the rim. While this assumption must be questioned in a more refined treatment, we show below that it leads to encouraging agreement with observations. Figure 5(a,b,c,d) shows model light curves obtained under these assumptions for the acceleration function Eq. (\ref{acc1}), using sample parameters $(a_0,a_1,h_0) = (5,0,0)$, $(0,5,0)$, $(0,20,1)$, and $(50,0,2)$, which we denote as models A, B, C, and D respectively. Table 1 lists additional simulation parameters which are common to all these models. In models A-C a common weight factor of 5 was assigned to the outermost ring of primary electrons to represent the excess rim current, while a factor 3 was used for model D. (The simulation would assign a weight factor of 1 to this ring for a uniform distribution.) In each case the rim weights were chosen to obtain reasonable fits to the observed Vela light curve. Since each simulation includes a total of 10 concentric primary current rings covering the PC interior, these rim weight factors increase the total PC currents above their uniform component values by factors of roughly 1.7 for models A-C and 1.4 for model D. For comparison, in each of these plots the light curves which would be produced by the uniform current alone (without the excess rim component) are shown in gray. If we compare these model results with the observed Vela light curve (Kanbach et. al 1994) shown in Figure 9, we see that the acceleration parameters which best match the observations are those for which the acceleration near the surface is low. In fact, satisfactory fits are obtained only if the primaries do not reach $\gamma \gtrsim 10^6$ until after they have attained heights $h \gtrsim 1$. If they exceed these energies at altitudes too far below the cutoff height $h_c$, the total cascade emission which they produce over the full acceleration region and beyond is spread over large solid angles, yielding broad pulse peaks. In particular, this tendency rules out constant-acceleration models ($a_1 = 0$) such as that shown in Figure 5(a), except in cases where $h_0 \gtrsim 2$ as in Figure 5(d). A comparison of Figures 5(b), 5(c) shows that even for linear acceleration ($a_0 = 0$), the fits are significantly improved by introducing nonzero values of $h_0$. Among the sample runs shown in Figure 5, models C and D show peak duty cycles which are in the best agreement with the observed values. Moreover, in each of these cases the first half of the interpeak emission resembles both the magnitude and slope seen in the data. This example shows that the two-component model for the primary current can produce consistent agreement with a significant portion of the total light curve. Unfortunately the agreement breaks down for the trailing interpeak component, but since our model assumes axisymmetric current rings it cannot account for any strong asymmetry in the light curve. Finally we note that all these models predict a low but finite level of emission throughout the phase interval between Peak 2 and Peak 1 (i.e., over regions outside the PC rim). This emission is due to the residual, high-altitude cascades which we suggested in DH94 might be the source of the interpeak emission. Kanbach et al. (1994) find no detectable emission in this phase interval for Vela, and no evidence for unpulsed emission. Given their stated estimates for the EGRET detector sensitivity, however, their findings are not in conflict with our model results for the sample datasets C and D described above. However, the observations do impose an additional constraint on the relative weight factors for the two-component PC current distribution. For example a uniform PC current, without any rim current enhancement, would produce significantly more emission outside the peaks than the observations allow. \section{PHASE-RESOLVED ENERGY SPECTRA} The same choices of parameters (model C and D) which best match the observed light curves in Section 6 also produce the best fits for the energy spectra. In spite of the similar appearance of their light curves, however, model D produces better spectral fits than model C. In fact, as shown in Figure 6 model D provides the closest match to the observed total (phase-averaged) spectrum across five decades in energy, The spectral differences among these models are principally due to their varying extent of cascade development. In models A and B the primary electrons reach maximum energies of $7.5 \times 10^6$ and $1.2 \times 10^7$ respectively, compared to $1.7 \times 10^7$ for Model C and $2.0 \times 10^7$ for Model D. The values for models A and B especially are too low to supply either the photons up to 3 GeV or the level of emission observed below 100 MeV. Clearly, the observed Vela total emission is not the result of curvature radiation alone. Figure 7 shows that this agreement for model D applies not only to the total spectra, but also to the phase-resolved spectra observed by EGRET (Kanbach et. al. 1994). These plots show fits for various phase intervals defined by these authors in their power law-fits to the Vela phase-resolved spectra for energies between 70 and 4000 MeV. The normalization factors were determined separately at each phase interval to match the data and differ by less than a factor of 2. We note that the model reproduces the tendency for the (quasi) power-law spectra at the phase intervals of the two peaks to become significantly softer than the spectra for the interpeak subintervals. In the Interior Scenario (Section 5) this trend is expected since the interpeak emission is due to the interior primary electron current, whose hard CR emission is less efficiently converted to softer cascade photons (cf. Figure 3). The IS model also reproduces the observational feature that the high-energy turnovers in the Vela spectra occur at lower energies for the peaks vs. the interpeaks. The sharpness of the high-energy turnovers in the P1 and P2 spectral intervals, due to magnetic one-photon pair production attenuation, are also reproduced, especially in P1. The model D spectra in the phase intervals LW1 and TW2, the emission just outside the peaks, turnover more gradually and at energies below 500 MeV. This emission is primarily curvature radiation at high altitudes from primary electrons that have lost a significant amount of their maximum energy. These phase intervals are thus predicted to have the softest spectra, consistent with both the data and the high indices of the power law fits of Kanbach et al. (1994). In model D, the hard spectra in intervals I1 and I2 extend to energies below 10 MeV, predicting that the interpeak emission should decrease relative to that of the peak emission at lower energies. This appears to be verified by the 0.07 - 0.6 MeV light curves measured by OSSE (Strickman et al. 1995), where no interpeak emission was detected. One quantitative measure of the spectral evolution during each pulse is the hardness ratio $H$, defined here as the ratio of the flux over 300 MeV to the flux between 100 and 300 MeV. Figure 8 shows the model D hardness ratio vs. pulse phase for $\zeta = 16^{\deg}$, corresponding to the phase-resolved spectra in Figure 7. The trend toward harder spectra during the interpeak phase interval is clear and appears to be consistent with EGRET Vela observations (Fierro et al. 1995). \newpage \section{TOTAL GAMMA FLUX ESTIMATES} If we identify the uniform component of our model PC current with the GJ current predicted by Eq. (\ref {gjc}), we can estimate an upper limit on the absolute $\gamma$-ray flux levels expected from our model sources within any specified energy range $\Delta E_\gamma$. The required inputs are the dataset sky map counts, the pulse period $P$, and the estimated distance $D$ (which we take to be 500 pc for Vela). We outline the procedure briefly as follows. First we derive the effective number of primaries traced in the simulation, taking into account the weight factors assigned to each concentric ring of electrons. Following the arguments in Section 5, we resolve this total number of primaries into two components representing uniform and rim distributions respectively. As noted above, our total flux estimate (uniform plus rim components) assumes that the uniform component is a GJ current. For the flux estimate, the quantity of interest is the number of GJ primaries in the simulation. After summing the full 3D photon arrays over the energy range $\Delta E_{\gamma}$ to produce the appropriate 2D sky maps, we find the number $\Delta N_\gamma$ of photons accumulated along a 1-bin strip of constant $\zeta$ and angular width $d \zeta$ during one full pulse ($\Delta \phi = 2 \pi$). The phase-averaged $\gamma$-ray flux $F_\gamma$ {\it per primary electron\/} at the distance D is then given by \begin{equation} \label{flx} F_\gamma = \Delta N_\gamma / 2 \pi \sin \zeta d \zeta P D^2 N_{GJ} \end{equation} where $N_{GJ}$ denotes the effective number of GJ primaries in the dataset (excluding the excess rim component). Finally we obtain an absolute total flux estimate by multiplying Eq. (\ref{flx}) by the (maximal) current of GJ primaries from the PC surface as given by Eq. (\ref{gjc}). The predicted fluxes for our Vela models A,B,C,D at energies $ > 100$ MeV as found from this procedure are $7.3 \times 10^{-5}$, $1.6 \times 10^{-4}$, $2.8 \times 10^{-4}$, $2.8 \times 10^{-4}$ photons ${\rm cm}^{-2} {\rm s}^{-1}$ respectively. It turns out that these values are all an order of magnitude higher than the average flux observed by EGRET (Kanbach et al. 1994), namely $(7.8\pm 1.0) \times 10^{-6}$ photons ${\rm cm}^{-2} {\rm s}^{-1}$ for $E_{\gamma} > 100$ MeV. Our high model flux levels, which obviously are due to strong beaming factors of the hollow-cone emission, are not by themselves a problem for our model since the GJ estimate should properly be regarded only as an upper limit on the PC current. We note, however, that the model flux estimate does fall closer to the GJ limit as the $\gamma$-beam half-angle $\theta_b$ is increased. In this respect the excess predicted flux shows that even larger PC dimensions and/or acceleration cutoff heights can be allowed within the framework of the model. \section{COMPARISON WITH OBSERVATIONS AT OTHER WAVELENGTHS} In the preceding sections we have applied the SPC model specifically to the Vela pulsar, in part because both the $\gamma$-ray light curves and phase-resolved spectra for this object have been observed in considerable detail. However, our model results for Vela can also account in general terms for the $\gamma$-ray emission from other pulsars with doubly-peaked profiles such as the Crab, Geminga, and PSR B1951+32 (Ramanamurthy et al. 1995). The second general class of light curves predicted by the SPC model, namely those with only a single broad peak, may describe PSR B1055-52 (but see below). At present the only source whose $\gamma$-ray light curve may be difficult to accomodate is PSR B1706-44 (Thompson et al. 1992), since recent EGRET observations (Thompson et al. 1995) suggest that this object may have a triply-peaked pulse. However, we must consider whether the SPC $\gamma$-ray model is also compatible with observations of pulsed emission at other wavelengths from Vela and the other known $\gamma$-ray pulsars. Our primary concern here involves the possible implications of these observations regarding the viewing geometry for each source. In this context we focus especially on three $\gamma$-ray pulsars for which we also have strong evidence of thermal X-ray emission from the NS surface, namely Vela itself (\"Ogelman et al. 1993), Geminga (Halpern and Holt 1993), and PSR B1055-52 (\"Ogelman and Finley 1993). These objects are of particular interest since the modulation and phase behavior of the X-ray emission should be directly related to the magnetic field geometry at the NS surface. To facilitate the discussion of these sources, in Figure 9 we have assembled their light curves at various wavelengths using a common phase origin for each source. It turns out that each object presents a distinct set of challenges for our model, which we analyze separately below. Although it shows no evidence of surface thermal X-ray emission we must also consider observations at other wavelengths from the Crab pulsar. The Crab has the distinction of having doubly peaked light curves in phase at all observed wavelengths. However, its optical emission exhibits polarization swings which cause special problems for the SPC model. An additional challenge is presented by recent HST and ROSAT imaging of the inner Crab nebula, which strongly suggest an observer angle $\zeta \lesssim 60^{\deg}$ (Hester et al. 1995). (a) Vela (PSR B0833-45) As shown in Figure 9(a), the pulsed radio emission from Vela (see for example Manchester and Taylor 1977) exhibits a single narrow peak which leads the first $\gamma$-ray peak by $\sim 0.12$ in phase (Kanbach et al. 1994). The radio pulse shows a high degree of linear polarization with an unusually wide swing ($\gtrsim 90^{\deg}$) in the polarization angle $\psi$ across the pulse. This behavior has been interpreted (Radhakrishnan and Cooke 1969, see also Michel 1991) in terms of the rotating projection of a dipolar magnetic field in the plane orthogonal to the viewing direction. In this model $\psi$ is given as a function of $\alpha$, $\zeta$, and the pulse phase angle $\phi$ by \begin{equation} \label{pol} \tan \psi = \sin \alpha \sin \phi / (\sin \zeta \cos \alpha - \cos \zeta \sin \alpha \cos \phi) \end{equation} Several authors (e.g. Lyne and Manchester 1988, Rankin 1990) have attempted to invert this relation to determine the values of $\alpha$ and $\zeta$ for various pulsars, although the results to date are subject to controversy (Michel 1991, Miller and Hamilton 1993). However, Eq. (\ref{pol}) does imply that the maximum rate of the polarization swing $R \equiv |d(\tan \psi) / d(\sin \phi)|$ occurs at the phase corresponding to the closest approach of the magnetic axis to the observer direction, which we denote by $\phi_{M}$. If this model is correct, the rapid, extended swing for Vela ($R \sim 5.9$) indicates that the observer viewpoint approaches a magnetic pole to within a few degrees. As may be seen from Figure 10, the values of $\alpha$ and $\zeta$ used in the Vela model datasets discussed in Sections 6 and 7 do not produce polarization swings which are either as rapid or extended as the observed values. However, the real challenge in accounting for the radio pulse in our model is not simply to find better combinations of these parameters. The key point is that if the radio pulse does indeed mark the phase of closest approach to either of the magnetic poles, in the case of Vela its location relative to the $\gamma$-ray peaks is inconsistent with the SPC model. In particular, the Interior Scenario requires $\phi_{M}$ to lie midway between the two $\gamma$ peaks, whereas in the Exterior Scenario it is displaced from the midpoint by $0.5$ in phase. In contrast, Kanbach et al. (1994) find the phases of the $\gamma$ peaks (relative to the phase of the radio peak, $\phi_0 = 0$) to be $\phi_{p1} = 0.12$ and $\phi_{p2} = 0.54$ respectively. Hence the standard PC model of the radio pulse asserts that $\phi_{M} = \phi_0 = 0$, while the IS predicts $\phi_{M} = (\phi_{p1} + \phi_{p2})/2 = 0.33$ and the ES has $\phi_{M} = 0.83$. Thus the standard model of the Vela radio pulse is inconsistent with the SPC $\gamma$-ray model. On the other hand, it turns out that both the optical and X-ray light curves for Vela fit much more naturally within the geometry of the IS. As shown in Figure 9(a), the optical emission (Wallace et al. 1977) has a doubly peaked light curve with a smaller peak-to-peak phase separation ($\sim 0.2$) than that seen in the $\gamma$-ray regime. Moreover, the $\gamma$-ray peaks enclose the optical peaks in the sense that the leading optical peak follows the leading $\gamma$ peak, while the opposite occurs for the trailing peaks (see for example Manchester and Taylor 1977). In the IS, this sort of optical/$\gamma$ phase relationship would hold if the optical and $\gamma$ emission were beamed in coaxial hollow cones from the PC, with beam angles $\theta^{opt}_{b} < \theta^{\gamma}_{b}$. This in turn suggests that the optical emission might either be associated with interior PC currents, or that it might be produced by the rim current at lower heights than the $\gamma$-emission. Figure 9(a) also shows the pulsed X-ray emission from Vela detected by the ROSAT satellite (\"Ogelman et al. 1993), which consists of a broad pulse trailing the radio peak, with the bulk of the emission occurring between the two $\gamma$-ray peaks. The harmonic content of the pulse suggests a complex nonsinusoidal structure, although the available X-ray data do not show firm correlations with the optical or $\gamma$ peaks (or clear evidence of more than one peak). The statistics are unfortunately limited by the fact that the emission contains contributions from the compact nebula as well as the pulsar, and the pulsed fraction of the latter is only about 11\%. \"Ogelman et al. (1993) obtain their best fit to the pulsed component with a soft blackbody spectrum ($T_6 \sim 1.5-1.6$). They also note that the total point source (pulsed plus unpulsed) can either be fit with a blackbody spectrum at a similar temperature or with a steep power law ($\Gamma \sim -3.3$), compared to a harder power law ($\Gamma \sim 2.0$) which fits the surrounding compact nebula. \"Ogelman et al. (1993) suggest that if the pulsed component is actually thermal emission, the modulation may be due either to a nonuniform surface temperature distribution or to anisotropic radiation transfer effects in the magnetosphere. In either case the key point for our model is that the pulsed X-ray emission should then be concentrated near the phase $\phi_{M}$ of closest approach of the observer direction to a magnetic pole (Page 1995). To the extent that the bulk of the emission does occur between the $\gamma$ peaks, the Vela X-ray light curve appears compatible with the IS $\gamma$-ray model. In summary it appears that the observed optical, X-ray, and $\gamma$-ray light curves for Vela all seem mutually consistent with the IS, whereas the radio polarization swing cannot have the usual interpretation based on Eq. (\ref{pol}) in either the IS or the ES. At present we have no satisfactory way to account for the phase of the Vela radio pulse within the general framework of any SPC model, unless we invoke the possibility of nondipolar magnetic fields near the NS surface. However, we should point out that this incompatiblity is not simply a problem for our $\gamma$-ray model. The same conflict already exists between the standard radio model and the entire class of thermal X-ray models (e.g. Page 1995) in which the peak(s) in the pulsed emission coincide with the closest approach of the magnetic pole(s) to the observer viewpoint. (b) Geminga (PSR B0630+178) Although Geminga has long been known to be a strong $\gamma$-ray source (Kniffen et al. 1975), it was first discovered to be a pulsar from X-ray observations (Halpern and Holt 1992). Shortly thereafter $\gamma$-ray pulses were detected at the X-ray period (Bertsch et al. 1992). To date no pulsed emission has been found at either radio or optical wavelengths, although an optical counterpart has been identified (Bignami et al. 1993). While the lack of optical and radio light curves prevent the sort of phase comparisons we can make for other sources, both the X-ray and $\gamma$-ray data are relatively rich in detail. Figure 9(b) shows the light curves for Geminga at both hard and soft X-ray energies from ROSAT observations (Halpern and Ruderman 1993) as well as in the EGRET $\gamma$-ray regime (Mayer-Hasselwander et al. 1994, Ramanamurthy 1995). As in the case of Vela, the $\gamma$-ray light curve above 100 MeV exhibits a two-peak structure with significant interpeak (bridge) emission. The peaks have duty cycles only moderately larger than in Vela, with a phase separation of $0.5$. In contrast, Halpern and Ruderman (1993) find that the X-ray light curves at both soft (0.07-0.53 keV) and hard (0.53-1.50 keV) energies consist of broad single pulses. The hard component is somewhat narrower, but perhaps most remarkably the soft and hard components are $\sim 105^{\deg}$ out of phase. Halpern and Ruderman (1993) have fit the hard and soft components of the pulsed X-ray spectrum to two blackbody sources at temperatures $T_{6} \sim 0.5$ and $\sim 3$ respectively. These authors suggest that the soft emission is from the overall NS surface, while the hard component arises from hotter regions around a PC. However, they also note that within the available statistics a power-law fit for the harder component is nearly as good as the blackbody fit, which leaves open the possibility of magnetospheric emission mechanisms. In any event the hot PC model of the hard X-ray emission appears to be consistent with the SPC $\gamma$-ray model, as in the case of Vela, since as seen in Figure 9(b) the bulk of the hard X-ray pulse from Geminga also lies between the double $\gamma$-ray peaks (Halpern and Ruderman 1993). Unfortunately the modulation of the soft X-ray component and its phase shift relative to the hard component complicate this model. In fact the hard and soft components may not be consistently explained within the framework of {\it any\/} NS heating/cooling models which assume dipolar magnetic-field symmetry. This point has led Halpern and Ruderman (1993) to suggest an off-axis dipole model in the case of Geminga. (c) PSR B1055-52 This source has been detected by EGRET at energies above 300 MeV (Fierro et al. 1993). Figure 9(c) shows that in contrast to the doubly peaked radio pulse, the $\gamma$-ray light curve appears to exhibit a single broad peak. However, the available statistics are insufficient to rule out a multipeaked substructure. The limited data makes it difficult to analyze the phase relationship between the radio and $\gamma$ pulses, although it may be significant that the precursor of the main radio pulse appears just at the trailing end of the $\gamma$ peak. It is noteworthy that the radio profile has some similarity to that of the Crab, including a peak-to-peak phase separation $\gtrsim 0.4$ which would require an off-axis dipole in an orthogonal rotator model. PSR 1055-52 has the distinction of exhibiting the hardest phase-averaged $\gamma$-spectrum of all the $\gamma$-ray pulsars known to date, with a photon spectrum index of $\sim 1.2$. It is worth noting here that PC cascades can definitely exhibit such hard spectra, although they tend to do so only when both the electron CR losses and pair-conversion rates are comparatively low. These conditions are most likely to apply in specific regions of the magnetosphere, especially close to the magnetic axes and/or at heights of several NS radii above the surface. However, both more detailed $\gamma$-ray observations and further modeling of this source will be required to determine how the hardness of the spectrum may constrain the SPC model. Pulsed X-rays have also been detected from PSR B1055-52 by ROSAT (\"Ogelman and Finley 1993). As in the case of Geminga, the emission exhibits distinct hard and soft components above and below $\sim 0.5$ keV, both of which exhibit broad single pulses. Figure 9(c) shows the phase relationships between the X-ray light curves and the pulses at radio and $\gamma$-ray energies. As in the $\gamma$-ray regime, evidence for substructure in either X-ray component is limited by the available statistics. Another striking similarity with Geminga is the large relative phase shift between the hard and soft X-ray peaks, with the hard component in this case leading by $\sim 120^{\deg}$. \"Ogelman and Finley (1993) obtain satisfactory spectral fits using two-component blackbody models, although they find that the hard component may also be fit by a power law which extrapolates up to flux levels in the $\gamma$-ray regime comparable with the EGRET observations. If PSR B1055-52 does in fact have only one $\gamma$-ray peak, then its relationship to the X-ray emission may be difficult to explain within the SPC model. The key problem is that the model identifies the phase of a single $\gamma$ peak with the phase $\phi_M$ of closest approach of the PC. However, if the hard X-ray component is due to PC heating as proposed for Geminga (Halpern and Ruderman 1993), the X-ray peak indicates a value for $\phi_M$ in apparent conflict with the $\gamma$-ray location. While this difficulty does not arise if the hard X-rays have a magnetospheric origin as \"Ogelman and Finley (1993) suggest, their phase shift relative to the $\gamma$-ray pulse is still problematical. As in the case of Geminga, however, the modulation of the soft X-ray component and its phase shift relative to the hard component complicate the picture. The fact that the radio pulse for PSR B1055-52 has two peaks, with noteworthy similarities to the Crab radio profile, is also puzzling. However, the principal question regarding the viability of the SPC model for this source is whether the $\gamma$-ray light curve is singly peaked. Hopefully further analysis of EGRET data will be able to resolve this question. (d) The Crab Pulsar (PSR B0531+21) In constrast to all other $\gamma$-ray pulsars, the light curve of the Crab exhibits a doubly peaked structure at at all wavelengths observed to date, with the peaks appearing at essentially the same phase positions throughout the entire spectrum. In purely geometric terms this phase synchronization seems to suggest that a variety of emission processes, which may occur in distinct magnetospheric regions of other pulsars, are spatially coincident in the Crab. In the context of SPC $\gamma$-ray models it appears to motivate a search for radio, optical, and X-ray emission mechanisms involving the cascade pairs. Unfortunately, this approach leads to at least one serious difficulty for the SPC model, namely the optical polarization swings found to occur across each peak (Smith et al. 1988). If both the optical and $\gamma$ peaks do originate from the same PC rim regions, then the optical swings cannot be due to the sort of rotational projection effect described by Eq. (\ref{pol}) since the extent of the swing through the phase intervals containing each $\gamma$-peak cannot exceed a few degrees (cf. Figure 10). However, SPC models for the Crab appear to be compatible in this respect with the radio pulses, which do not exhibit significant polarization swings. In addition to this problem, a significant constraint on SPC models of the Crab pulsar is posed by recent HST and ROSAT observations of the inner nebula (Hester et al. 1995). These observations appear to confirm numerous earlier suggestions that the observer angle $\zeta$ for the Crab is considerably larger ($\lesssim 60\deg$) than the values ($\sim 15\deg$) used in our sample Vela datasets. However, this finding does not by itself rule out the SPC model for the Crab, since it turns out that such large values of $\zeta$ can be accomodated if we allow the PC dimensions to be $\sim 4 \-- 5$ times larger than the standard estimate Eq. (\ref{tpc}), as opposed to the factor $2$ used in our model datasets for Vela. Somewhat smaller values are also adequate if the cascades are assumed to extend up to heights $\gtrsim 3$ NS radii. Thus in the case of the Crab especially, the dimensions of the PC are critical to our model. \newpage \section{DISCUSSION} The model we have presented here has at least one significant advantage over an alternative SPC model (Sturner and Dermer 1994, Sturner et al. 1995) in which PC cascades are initiated by Comptonizaton of primaries by soft photons from the NS surface rather than CR emission. As we have shown, extended primary acceleration can easily generate CR-induced cascades at heights reaching up to several NS radii. In contrast, cascades due to Comptonization should be confined to significantly lower regions unless some mechanism for strong beaming of the soft photons is invoked. Assuming that similar PC dimensions are used in both models, the Comptonization model has a more limited ability to overcome the observability problem. The best results we have obtained to date from the extended cascade SPC model are for those cases in which the net electron acceleration becomes significant only at heights $h \gtrsim R_{ns}$ above the NS surface. However, we have shown in Section 3 that neither resonant Compton scattering of thermal photons from the NS surface nor other known energy loss processes considered in previous PC models can effectively counteract accelerating potentials of the types we have considered over distances of this order. This applies in particular to resonant Compton scattering, even if we assume the highest plausible values for both the surface temperatures and thermal PC radii. Thus it is obviously important to investigate the possibilility noted in Section 5, namely that downward-oriented cascades initiated by reversed secondary acceleration can prevent or impede acceleration just above the surface. An obvious next step in the exploration of the SPC model is to trace the development of downward-oriented cascades in detail, and if possible to estimate both their significance as a source of energetic Comptonizing photons and the depth of the surface plasma layer which they may create. The discussion in Section 9 shows that the phase relationships between light curves at different wavelengths are in fact quite complex. The problem of accounting for all these observations in a self-consistent manner may eventually force us to consider models with asymmetric magnetic field geometries. One initial step in this direction would be to consider off-axis dipolar models of the type suggested by Halpern and Ruderman (1993) in more detail. In such models we anticipate that the modulation of thermal X-ray emission from, say, the PC surface may be significantly out of phase with magnetospheric emission produced above the surface and directed along the open field lines. \section{ACKNOWLEDGEMENTS} We are indebted to Joe Fierro, Gottfried Kanbach, Peter Michelson, P.V. Ramanamurthy, and David Thompson for valuable discussions regarding EGRET observations, and to Mark Strickman for information regarding OSSE and COMPTEL results. We also thank Hakki \"Ogelman and John Finley for discussions on the pulsed X-ray emission from Vela. Michal Marko provided valuable assistance in the development of our visualization and analysis software. We gratefully acknowledge support for this work from NASA CGRO Guest Investigator Grants for Phases 3 and 4 (JKD, AKH), and from the NASA Astrophysics Theory Program (AKH). \newpage
\section{Introduction} Observations of anisotropy in the Cosmic Microwave Background Radiation (CMBR) yield valuable clues about the formation of large-scale structure in the early universe. A particularly interesting angular scale for observing CMBR anisotropy is near 0\fdg5, where the first ``Doppler peak'' (or adiabatic peak) enhancement of the fluctuation power spectrum is expected to be observable (\cite{white94}). The Medium Scale Anisotropy Measurement (MSAM) is an experiment designed to measure CMBR anisotropy at this angular scale. This paper reports the initial results from the second flight of this experiment. A number of detections of anisotropy at angular scales near 0\fdg5 have been reported recently. Observations by ARGO (\cite{debernardis94}), the Python experiment (\cite{dragovan94}), the fourth flight of the MAX experiment (\cite{devlin94,clapp94}), SK94 (\cite{netterfield94}), and SP94 (\cite{gundersen94}) all report detections of anisotropy near this angular scale. Quantifying CMBR anisotropy at the level of these detections is an extremely challenging observational task (\cite{wilkinson95}). Many potential systematic errors cannot be unequivocally ruled out at the necessary levels, with the result that any single observation cannot prudently be accepted without an independent confirmation. The results in this paper are our attempt to confirm the results of our previous work. By observing the same region of the sky with a second balloon flight, we demonstrate the repeatability of our measurements in the presence of potential atmospheric noise and contamination from Earthshine. We have reported earlier (\cite{cheng94}, hereafter Paper~I) our observations of anisotropy of the CMBR from the first flight of MSAM in 1992. Our results from those observations were 1) a positive detection of anisotropy, with the caveat that we could not rule out foreground contamination by bremsstrahlung; 2) the identification of two particular bright spots that were consistent with being unresolved sources. This paper reports our first results from the 1994 flight of MSAM, which observed an overlapping field. \section{Instrument Description} This instrument has been briefly described in Paper~I; we give only an overview here. It has four spectral bands at 5.6, 9.0, 16.5, and 22.5~${\rm cm}^{-1}$, giving sensitivity to CMBR and Galactic dust. The off-axis Cassegrain telescope forms a 30\arcmin\ beam on the sky. The chopping secondary mirror moves this beam in a step motion 40\arcmin\ left and right of center. The beam moves center, left, center, right with a period of 0.5~s. The detectors are sampled at 32~Hz, synchronously with the chop. The telescope is mounted on a stabilized balloon-borne platform. The absolute pointing reference is provided by a star camera; positions between camera fixes are interpolated using a gyroscope. The telescope is shielded with aluminized panels so that the dewar feed horn, the secondary and most of the primary have no direct view of the Earth. The gondola superstructure was changed between the 1992 and 1994 flights. The previous superstructure as viewed from the telescope had a substantial cross-section of reflective material; in spite of our efforts to shield it we were concerned about the telescope being illuminated by reflected Earthshine. The new design is a cable suspension with considerably lower cross section above the telescope. Ground measurements indicate that rejection of signals from sources near the horizon is better than 75~dB in our longest wavelength channel. \section{Observations} The package was launched from Palestine, Texas at 00:59~UT 2~June~1994, and reached its float altitude of 39.5~km at about 03:25~UT. Science observations ended with sunrise on the package at 12:04~UT. During the flight we observed Jupiter to calibrate the instrument and map the telescope beam, scanned M31 (which will be reported in a future {\sl Letter\/}), and integrated on the same CMBR field observed during the 1992 flight for 3.5 hours. The CMBR observations were made as described in Paper~I. The telescope observes near the meridian 8\arcdeg\ above the north celestial pole, and scans in azimuth $\pm 45\arcmin$ with a period of 1 minute. The scan is initially centered on a point 21\arcmin\ to the east of meridian. We track to keep this point centered in our scan until it is 21\arcmin\ to the west of meridian, then jog 42\arcmin\ to the east. Each scan takes about 20 minutes, and half of each scan overlaps the preceding scan. We completed 4.5 such scans from 05:12 to 06:38~UT (we call this section 1 of the data), and completed an additional 7 scans from 07:22 to 09:43~UT (section 2). The observed field is two strips at declination $81\fdg8 \pm 0\fdg1$, from right ascension 15\fh27 to 16\fh84, and from 17\fh57 to 19\fh71 (all coordinates are J1994.5). Fig.~\ref{f_fields} shows the fields observed in the 1992 and 1994 flights. The overlap between the fields is better than half a beamwidth throughout the flight. Our ability to observe exactly the same position on the sky is currently limited by the error in determining the position of the IR beam center during the initial in-flight calibration, i.e., our real-time determination of pointing is not as accurate as our post-flight determination. \section{Data Analysis} The signal from the detectors is contaminated by spikes induced by cosmic rays striking the detectors; we remove these spikes. The data are calibrated by our observation of Jupiter. The absolute pointing is determined from star camera images. The detector data are analyzed to provide measurements of brightness in our four spectral channels as a function of sky position. These are then fit to a spectral model to produce measurements of CMBR anisotropy and dust optical depth. These analyses and their results are described in the following sections. \subsection{Pointing} We determined the pointing by matching star camera images against a star catalog. This fixes the position of the camera frame at the time the exposure was taken. Between exposures, position is interpolated with the gyroscope outputs plus a small linear correction to make the gyroscope readings consistent with the camera fixes. This correction is typically 2\arcmin\ in 20~minutes. The relative orientation of the camera frame and the IR telescope beam is fixed by a simultaneous observation of Jupiter with the camera and the IR telescope. The resulting absolute pointing is accurate to 2\farcm5, limited by the gyroscope drift correction. The pointing analysis was done in an identical way for the 1992 flight, and has similar accuracy. \subsection{Detector Data Reduction} The instrument is calibrated by in-flight observations of Jupiter. The brightness temperatures of Jupiter for our four spectral channels are 172, 170, 148, and 148~K, derived from the spectrum of Jupiter observed by \cite{griffin86}. The apparent diameter of Jupiter during the 1994 flight is 42\arcsec. The uncertainty in the absolute calibration is 10\%, dominated by uncertainty in the antenna temperature of Jupiter. The relative calibration uncertainty between the 1992 and 1994 flights is 5\%, due to noise in the observations of Jupiter. The detector signal contains spikes, at a rate of 0.25--0.5~${\rm s}^{-1}$, consistent with the hypothesis that they are due to cosmic rays striking the detectors (\cite{charakhchyan78}), and with the rate reported in Paper~I. Cosmic rays deliver an unresolved energy impulse to the detector; we remove them by fitting the data to the impulse response function of the detector/amplifier/filter chain. We give here our results for the 5.6~${\rm cm}^{-1}$\ channel; the numbers for the other channels are similar. Candidate spike locations are identified using a $1.5\,\sigma$ threshold. The data within 1~s (5 detector time constants) are fit to a model of the response function. About 2\% of the spikes require a second spike 2--10 samples separated from the first to be added to the fit. If the resulting spike amplitude has less than $3\,\sigma$ significance, the data is left as-is. If the fit is good, and the spike amplitude has more than $3\,\sigma$ significance, the spike template is subtracted. 5065 spikes are subtracted out of 504,000 time samples. (We allow either positive or negative amplitudes; 90\% of the spikes have positive amplitude.) If the fit is poor, and the spike amplitude is significant, full data records (64 samples, or 2 sec) before and after the spike are deleted. 317 spikes were eliminated this way, removing a total of about 6\% of the data. We estimate the instrument noise by measuring the variance in the demodulated, deglitched data after removing a slow drift in time and the mean in each sky bin. This estimate is made for each 20~minute segment of data, and is then propagated throught the remaining processing. All $\chi^2$ reported below are with respect to this error estimate. We divide the sky into bins that are small compared to the beamsize. The bins are 0\fh057 in right ascension and 0\fdg12 in declination. Due to sky rotation, the data also need to be divided by angular orientation of the beam throw on the sky; the bin size for this coordinate is 10\arcdeg. The data are then fit to a signal in each sky bin plus a model of long-term drift formed from a cubic spline with knots every 12 minutes (2.5 minutes for the 16.5~${\rm cm}^{-1}$\ channel), plus terms for gondola inclination, roll, and air pressure. The simultaneous fit of long-term drift and sky signal ensures that this fit does not bias our observations of the sky. This fit is done separately on each channel and section of the flight. The resulting sky signals have bin-to-bin correlation, and we propagate a full covariance matrix through the remainder of the analysis. Sky bins containing less than 4~s of integration are deleted. So that our error estimate, described in the preceding paragraph, is unbiased by sky signal, we form the estimate from the residuals of this fit, and iterate to obtain a consistent solution. The data are demodulated in two different ways. The double difference demodulation corresponds to summing the periods when the secondary is in the central position, and subtracting the periods when it is to either side. This demodulation is least sensitive to atmospheric gradients and gondola swinging. The single difference demodulation is formed by differencing the period when the secondary is to the right from that when it is to the left, and ignoring the periods when the secondary is in the center. We use the scan over Jupiter to deduce optimal demodulations of the infrared signal. The binned dataset contains 90\% of all the data originally taken, with an achieved sensitivity in each of the four channels of 240, 150, 80, and 230~\hbox{$\mu$K}~$\sqrt{\rm s}$ Rayleigh-Jeans. For channels 1 and 2 this is 490 and 850~\hbox{$\mu$K}~$\sqrt{\rm s}$ CMBR. The offsets in the demodulated data for the different channels and demodulations range from 1 to 6~mK~RJ, smaller than those reported in Paper~I. \subsection{Spectral Decomposition} At each sky bin, we fit the four spectral channels to a model consisting of a CMBR anisotropy plus emission from warm Galactic dust. The results are not very sensitive to the parameters of the dust model; we use a dust temperature of 20~K and an emissivity index of 1.5 (consistent with \cite{wright91}). The fit is done separately for the single and double difference demodulations. The $\chi^2/$DOF for the fit is 408/430 (double difference) and 448/430 (single difference). Fig.~\ref{f_dust} shows the resulting fitted dust optical depth at 22.5~${\rm cm}^{-1}$. For clarity this figure has been binned more coarsely and does not distinguish between points at slightly different declination or chop orientation; our analyses, however, do not ignore these details. We have fit our observations to the {\sl IRAS}\ Sky Survey Atlas at 100~\micron\ (\cite{wheelock93}) convolved with our beam patterns, with amplitude and offset as free parameters. The resulting fit is superimposed on Fig.~\ref{f_dust}. The $\chi^2/$DOF of this fit is 262/210 for the double difference demodulation and 310/210 for the single difference. The ratio of optical depths between IRAS and our data is consistent with an average dust emissivity spectral index between our bands and 100~\micron\ of $\alpha = 1.40 \pm 0.16$ (still assuming a dust temperature of 20~K). Our measurements of CMBR anisotropy are plotted in Fig.~\ref{f_cmbr}. Superimposed are the measurements from 1992. As noted earlier, there is non-negligible correlation between the error bars on different sky bins. In making Fig.~\ref{f_cmbr} we have fit out the two largest eigenmodes of the covariance matrix, and used error bars formed from the diagonal of the covariance matrix after removing the two largest eigenmodes; the result is that the error bars shown in the figure can be approximately treated as uncorrelated. (This procedure is similar to that used in \cite{fixsen94b} for the {\sl COBE\/}/FIRAS calibration.) The data have also been binned more coarsely, as in Fig.~\ref{f_dust}. We stress that these steps are taken only for producing representative figures; in all quantitative analyses we use the full dataset and the full covariance matrix. We are in the process of calculating the correlation for the MSAM1-92 data; the 1992 data plotted here are identical to those in Paper~I. \subsection{CMBR Anisotropy} To set limits on anisotropy in the CMBR, we assume Gaussian fluctuations with a Gaussian-shaped correlation function. We set 95\% confidence level upper and lower bounds on the total rms fluctuation over the sky $(\sqrt{C_0})$, assuming this correlation function with a given correlation angle $\theta_c$. The method used is described in Paper~I, though we now use a full covariance matrix for the instrument noise on the observations. The upper and lower bounds from these observations for the single and double difference demodulations are shown in Fig.~\ref{f_deltat}. The bounds for the correlation angles at which the two demodulations are most sensitive are summarized in Table~\ref{t_deltat}, which also shows results for the two sections of the flight separately. The confidence intervals for both demodulations are consistent with those in Paper~I. \section{Conclusions} We observed the same field in our 1992 and 1994 flights in order to determine if the detected signal was due to sidelobe pickup, atmospheric noise, or other systematic effects, or was in fact present in the sky. While we are still in the process of completing a detailed quantitative comparison of the two datasets, it is apparent that the double difference CMBR anisotropy features reproduce quite well. This encourages us to believe that the signal we see in the double difference is present on the sky, and that contamination from atmosphere or sidelobes is small compared to the sky signal. The single difference CMBR signal does not appear to reproduce as well. Pending the completion of the more thorough comparison, we cannot rule out contamination in the single difference channel. In Paper~I we pointed out that the anisotropy we observe could be due to diffuse Galactic bremsstrahlung. This possibility remains, and will be addressed by our MSAM2 experiment, which will observe the same fields in five bands over 65--170~GHz. In Paper~I we raised the possibility that the ``sources'' at R.A. 19~h and 15~h were either foreground sources of a previously unknown population, or non-Gaussian CMBR fluctuations. This speculation was prompted by our belief that such features were inconsistent with Gaussian statistics. More careful analysis by us and independently by \cite{kogut94} has indicated that features like these are in fact consistent with a variety of plausible correlation functions. Observations by \cite{church95} at 4.7~${\rm cm}^{-1}$\ rule out the source MSAM15$+$82 being more compact than 2\arcmin. Therefore removal of these regions in studies of CMBR anisotropy, as we recommended in Paper~I, are a biased edit of the data, and we no longer recommend it. Our current conclusion is that the double difference, whole flight numbers in Table~\ref{t_deltat} are a reliable estimate of CMBR anisotropy in the observed regions. When we include the 10\% uncertainty in the calibration, the resulting limits are $\Delta T/T = 1.9^{+1.3}_{-0.7}\times 10^{-5}$ (90\% confidence interval) for total rms fluctuations. In the band power estimation of (\cite{bond95}), this is $\langle {\cal C}_l \rangle_B = 2.1^{+1.5}_{-0.9}\times 10^{-10}$ ($1\,\sigma$ limits), with $\langle l \rangle = 263$. The CMBR anisotropy channel, Galactic dust channel, pointing, covariance matrices, and beammaps are publicly available. For more information, read {\tt ftp://cobi.gsfc.nasa.gov/pub/data/msam-jun94/README.tex}. \acknowledgments We would like to thank the staff of the National Scientific Balloon Facility, who remain our willing partners in taking the calculated risks that result in extremely successful flights. W.~Folz and J.~Jewell traveled with us to the NSBF to help with flight preparations. T.~Chen assisted in building and testing our new star camera system. We are grateful to M.~Devlin and S.~Tanaka for providing cappuccino at the crucial moment in Palestine. The Free Software Foundation provided the cross-development system for one of the flight computers. This research was supported by the NASA Office of Space Science, Astrophysics Division. \clearpage \begin{deluxetable}{rccrrcrr} \tablecolumns{8} \tablecaption{Upper and lower bounds on total rms CMBR anisotropy ($\protect\sqrt{C_0}$) \label{t_deltat} } \tablehead{ \colhead{} & \colhead{} & \multicolumn{3}{c}{MSAM1-94} & \multicolumn{3}{c}{MSAM1-92} \\ \cline{3-5}\cline{6-8} \colhead{} & \colhead{} & \colhead{} & \colhead{Upper} & \colhead{Lower} & \colhead{} & \colhead{Upper} & \colhead{Lower} \\ \colhead{$\theta_c$} & \colhead{Section} & \colhead{R.A.} & \colhead{Bound} & \colhead{Bound} & \colhead{R.A.} & \colhead{Bound} & \colhead{Bound} \\ \colhead{} & \colhead{} & \colhead{(h)} & \colhead{(\hbox{$\mu$K})} & \colhead{(\hbox{$\mu$K})} & \colhead{(h)} & \colhead{(\hbox{$\mu$K})} & \colhead{(\hbox{$\mu$K})} } \startdata \cutinhead{Single Difference} 0\fdg5 & 1 & 15.27--16.84 & 163 & 40 \nl & 2 & 17.57--19.71 & 75 & 17 \nl & All & 15.27--19.71 & 79 & 30 & 14.44--20.33 & 116 & 53 \nl \cutinhead{Double Difference} 0\fdg3 & 1 & 15.27--16.84 & 132 & 44 \nl & 2 & 17.57--19.71 & 74 & 24 \nl & All & 15.27--19.71 & 78 & 34 & 14.44--20.33 & 97 & 50\nl \enddata \tablecomments{The limits in this table do not include the calibration uncertainty.} \end{deluxetable} \clearpage \bibliographystyle{aas}
\section{Introduction} % The four-dimensional self-dual Einstein equation (SdE) has been given attention for a long time both in physics and mathematics, as well as the self-dual Yang-Mills equation. Among a number of works associated with the SdE \cite{EGH}, an interesting and important subject is to connect it to other (possibly simple) field equations. Well-known examples of it are Plebanski's heavenly forms \cite{plebanski}, there the SdE is given in terms of one function of space-time coordinates. Q-Han Park \cite{park} and Ward \cite{ward} have shown that the SdE is derived from several two-dimensional sigma models with the gauge group of area preserving diffeomorphisms, SDiff(${\cal N}_2)$. Park also has clarified the correspondence between the sigma models and first and second heavenly forms. On the other hand, by Ashtekar's canonical formulation for general relativity \cite{ashtekar1}, the SdE has been reformulated as the Nahm equation \cite{ashtekar2}, and its covariant version is given in Ref.\,\cite{mason}. Through this formulation, Husain has arrived at one of sigma models, that is, the principal chiral model \cite{husain}. Also by several reduction methods, other interesting models, e.g. the SL$(\infty)$ (affine) Toda equation \cite{boyer}\cite{park}, the KP equation \cite{castro} etc., are obtained. \par Although we have various examples connected to the SdE, their relation is rather unclear since their derivations from the SdE are more or less complicated and separated. Such a link of the models, however, should be investigated in order to understand the SdE further and in particular to develop the quantization of self-dual gravity. \par In this paper, we describe the self-dual Einstein space by a trio of differential form equations for simple two-forms and derive several integrable theories quickly. This formulation elucidates their relation and may indicate the possibility to find further a large class of models connected to the SdE. \vskip 0.4cm \section{Self-dual Einstein equation}% We start from the observation that the SdE is expressed as closed-ness conditions of basis of the space of anti-self-dual two-forms, \begin{equation} d ( e^0 \wedge e^i - {1 \over 2} \epsilon_{ijk} e^j \wedge e^k ) = 0 \ , \qquad i = 1,2,3, \label{eq: cls} \end{equation} where $e^{0,i} = e^{0,i}_{\mu} dx^{\mu}$ are tetrad one-forms on four-manifold. This formulation was employed by Plebanski to reduce the SdE to the heavenly forms \cite{plebanski}, there the indices $i,j,k$ are replaced with spinor ones $A,B$ by the Pauli matrices ${(\sigma^i)_A}^B$. The equations in (\ref{eq: cls}) appear also in Ref.\,\cite{capovilla}\cite{abe}. We rewrite (\ref{eq: cls}) by defining a null basis, \begin{eqnarray} && {\cal Z} = e^0 + i e^1, \qquad \ {\bar {\cal Z}} = e^0 - i e^1, \nonumber \\ && {\cal \chi} = e^2 - i e^3, \qquad \ {\bar {\cal \chi}} = e^2 + i e^3 \ . \end{eqnarray} Then (\ref{eq: cls}) becomes \begin{equation} d ( {\cal Z} \wedge {\cal \chi}) = 0 \ , \qquad d ( {\bar {\cal Z}} \wedge {\bar {\cal \chi}}) = 0 \ , \qquad d ( {\cal Z} \wedge {\bar {\cal Z}} + {\cal \chi} \wedge {\bar {\cal \chi}}) = 0 \ . \label{eq: nc} \end{equation} \par As for symmetry, adding to the diffeomorphism invariance, (\ref{eq: nc}) is invariant under the local SL$(2,C)$ (self-dual) transformation, in matrix form, \begin{equation} \left[ \begin{array}{l} {\cal Z}^{'} \\ {\cal \chi}^{'} \end{array} \right] = \left[ \begin{array}{cc} a \ & b \ \\ c \ & d \ \end{array} \right] \left[ \begin{array}{l} {\cal Z} \\ {\cal \chi} \end{array} \right] \ , \qquad \left[ \begin{array}{l} {\bar {\cal Z}}^{'} \\ {\bar {\cal \chi}}^{'} \end{array} \right] = \left[ \begin{array}{cc} d \ & -c \ \\ -b \ & a \ \end{array} \right] \left[ \begin{array}{l} {\bar {\cal Z}} \\ {\bar {\cal \chi}} \end{array} \right] \ , \quad ad-bc=1 \ , \label{eq: sl} \end{equation} and also invariant under the global SL$(2,C)$ (anti-self-dual) transformation for the pairs $({\bar {\cal Z}}, {\cal \chi})$, $({\cal Z}, {\bar {\cal \chi}})$ with the same form as (\ref{eq: sl}). \par The two-form in the last equation in (\ref{eq: nc}) is of rank-four, while others are of rank-two, but we can re-express the last equation by defining two one-forms ${\cal P} = {\cal Z} - {\bar {\cal \chi}}$, ${\cal Q} = {\cal \chi} + {\bar {\cal Z}}$. Using them, we obtain the following equations equivalent to (\ref{eq: nc}), \begin{equation} d ( {\cal Z} \wedge {\cal \chi}) = 0 \ , \qquad d ( {\bar {\cal Z}} \wedge {\bar {\cal \chi}}) = 0 \ , \qquad d ( {\cal P} \wedge {\cal Q}) = 0 \ . \label{eq: smcl} \end{equation} Since all two-forms in (\ref{eq: smcl}) are simple and closed, they can be written as, on a local coordinate system, \begin{equation} {\cal Z} \wedge {\cal \chi} = dz \wedge dx \ , \qquad {\bar {\cal Z}} \wedge {\bar {\cal \chi}} = d\bar z \wedge d\bar x \ , \qquad {\cal P} \wedge {\cal Q} = dp \wedge dq \ , \label{eq: sm} \end{equation} where $(z,x,\bar z,\bar x,p,q)$ are functions. Although we use the notation suitable for the real, Euclidean case, we generally consider the complex SdE, so the bars in (\ref{eq: sm}) do not mean the complex conjugation in usual. {}From the definition of ${\cal P}$ and ${\cal Q}$, two identities are obtained, \begin{eqnarray} {\cal Z} \wedge {\cal \chi} \wedge {\cal P} \wedge {\cal Q} && = dz \wedge dx \wedge dp \wedge dq = dz \wedge dx \wedge d\bar z \wedge d\bar x = {\cal Z} \wedge {\cal \chi} \wedge {\bar {\cal Z}} \wedge {\bar {\cal \chi}} \ , \nonumber \\ {\bar {\cal Z}} \wedge {\bar {\cal \chi}} \wedge {\cal P} \wedge {\cal Q} && = d\bar z \wedge d\bar x \wedge dp \wedge dq = dz \wedge dx \wedge d\bar z \wedge d\bar x = {\cal Z} \wedge {\cal \chi} \wedge {\bar {\cal Z}} \wedge {\bar {\cal \chi}} \ . \label{eq: id} \end{eqnarray} (\ref{eq: smcl}), (\ref{eq: sm}) and (\ref{eq: id}) are key equations in our formulation. For later use, we define the notation $x^a = (z, \bar z)$ and $x^k = (p,q)$. \vskip 0.4cm \section{ Integrable theories derived from the SdE} % (a)${\it \, The \ principal \ chiral \ model}$ \\ At first, let us choose $(z,\bar z,p,q)$ as four coordinate variables and $(x,\bar x) = (A_z, A_{\bar z})$ as functions of them. Then (\ref{eq: id}) reads \begin{equation} dz \wedge dA_z \wedge dp \wedge dq = dz \wedge dA_z \wedge d\bar z \wedge dA_{\bar z} \ , \quad d\bar z \wedge dA_{\bar z} \wedge dp \wedge dq = dz \wedge dA_z \wedge d\bar z \wedge dA_{\bar z} \ , \label{eq: pcm0} \end{equation} from which we can quickly derive the following equations after expanding $dA_{z(\bar z)} = \partial_a A_{z(\bar z)} dx^a + \partial_k A_{z(\bar z)}dx^k$ and rescaling $A_{z(\bar z)}$ by $-2$, \begin{equation} \partial_z A_{\bar z} + \partial_{\bar z} A_z = 0 \ , \qquad F_{z\bar z}= \partial_z A_{\bar z} -\partial_{\bar z} A_z + \{A_z, A_{\bar z} \} = 0 \ , \label{eq: pcm} \end{equation} where $\{A_z, A_{\bar z} \}$ is the Poisson bracket with respect to $(p,q)$. Using the potentials $A_{z(\bar z)}$, we define generators ${\cal A}_{z(\bar z)} = \{\ \,, A_{z(\bar z)}\} = \partial_q A_{z(\bar z)} \partial_p - \partial_p A_{z(\bar z)} \partial_q$ of the algebra of area preserving diffeomorphisms sdiff(${\cal N}_2$), which, at each space-time point $(z, \bar z)$, act on function on the internal surface ${\cal N}_2$ parametrized by the coordinates $(p,q)$. The second equation in (\ref{eq: pcm}) means the generators ${\cal A}_{z(\bar z)}$ are pure-gauge, ${\cal A}_{z(\bar z)} = g^{-1} \partial_{z(\bar z)} g$, where $g$ is an element of the group SDiff(${\cal N}_2)$. Substituting them into the first equation, we obtain \begin{equation} \partial_{\bar z}(g^{-1}\partial_z g) + \partial_z (g^{-1}\partial_{\bar z} g) = 0 \ . \end{equation} This is precisely the chiral model equation on the $(z, \bar z)$ space-time with $(p,q)$ treated as coordinates on ${\cal N}_2$. Let us solve $A_{z(\bar z)}$ for a single scalar function $\Theta$ by the first equation in (\ref{eq: pcm}), that is, $A_z = 2 \partial_z \Theta$ and $A_{\bar z} = - 2 \partial_{\bar z} \Theta$. Then the second equation becomes \begin{equation} \Theta_{z,\bar z} + \Theta_{z,p} \Theta_{\bar z, q} - \Theta_{z,q} \Theta_{\bar z, p} = 0 \ , \end{equation} where $\Theta_{z, p} = \partial_z \partial_p \Theta$. By an adequate gauge condition for the local SL$(2,C)$ symmetry, tetrads can take the form, \begin{eqnarray} {\cal Z} = h^{1 \over 2} dz \ , \qquad && {\cal \chi} = - h^{1 \over 2} d\bar z - h^{-{1 \over 2}} \Theta_{z,k} dx^k \ , \nonumber \\ {\bar {\cal Z}} = h^{1 \over 2} d\bar z \ , \qquad && {\bar {\cal \chi}} = \ \, h^{1 \over 2} dz + h^{-{1 \over 2}} \Theta_{\bar z, k}dx^k \ , \qquad h = \{\Theta_{\bar z}, \Theta_z \} \ , \end{eqnarray} and the line element is \begin{equation} ds^2 = {\cal Z} \otimes {\bar {\cal Z}} + {\cal \chi} \otimes {\bar {\cal \chi}} = - \Theta_{a,k} \, dx^a \otimes dx^k + {1 \over \{\Theta_z, \Theta_{\bar z} \}} \Theta_{z,k} \Theta_{\bar z,l} \, dx^k \otimes dx^l \ . \end{equation} It is obvious that all self-dual metrices are obtained from this sigma model. In the case of $\{A_z,A_{\bar z}\}=0$, the volume form ${1 \over 4}({\cal Z} \wedge {\bar {\cal Z}} \wedge {\cal \chi} \wedge {\bar {\cal \chi}})$ vanishes, which corresponds to a degenerate space-time. \vskip 0.2cm (b)${\it \, The \ topological \ model \ with \ the \ WZ \ term \ only}$ \\ Next we take $(z,x,\bar z,\bar x)$ as our coordinates and $(p,q) = (B_0, B_1)$ as functions of them. After changing the notation $(z,x,\bar z,\bar x)=(z,\bar z,p,q)$, (\ref{eq: id}) gives \begin{equation} \{B_0, B_1 \}_{(z,\bar z)} = 1 \ , \qquad \{B_0, B_1 \} = 1 \ . \end{equation} The bracket in the first equation is defined with respect to $(z,\bar z)$. These equations are rather unfamiliar, but if we define $\partial_k A_{z(\bar z)} = \{B_0, B_1 \}_{(z(\bar z), x^k)}$, we can easily check the integrability $\partial_{[k} \partial_{l]} A_{z(\bar z)} = 0$ and \begin{equation} \partial_z A_{\bar z} - \partial_{\bar z} A_z = 0 \ , \qquad \{A_z, A_{\bar z} \} = 1 \ . \label{eq: tm2} \end{equation} Also in this case, generators ${\cal A}_{z(\bar z)} = \{\ \,, A_{z(\bar z)}\}$ of sdiff(${\cal N}_2$) are pure-gauge, ${\cal A}_{z(\bar z)} = g^{-1} \partial_{z(\bar z)} g$, and from the first equation, we obtain \begin{equation} \partial_{\bar z}(g^{-1}\partial_z g) - \partial_z (g^{-1}\partial_{\bar z} g) = 0 \ . \end{equation} This is the topological model derived from the lagrangian of the Wess-Zumino term only \cite{park}. The potentials $A_{z(\bar z)}$ are given in terms of one function $\Omega$ by the first equation in (\ref{eq: tm2}). Then $A_{z(\bar z)}= \partial_{z(\bar z)} \Omega$ and the second equation becomes \begin{equation} \Omega_{z, p} \Omega_{\bar z, q} - \Omega_{z, q} \Omega_{\bar z, p} = 1 \ , \label{eq: pl} \end{equation} which is Plebanski's first heavenly form \cite{plebanski}. With a gauge condition for the local SL$(2,C)$ symmetry, tetrads are given by \begin{equation} {\cal Z} = dz \ , \quad {\cal \chi} = d\bar z \ , \quad {\bar {\cal Z}} = \Omega_{z,k} dx^k \ , \quad {\bar {\cal \chi}} = \Omega_{\bar z,k} dx^k \ , \end{equation} and the line element is $ds^2 = \Omega_{a,k} dx^a \otimes dx^k$. \vskip 0.2cm (c) ${\it \, The \ WZW \ model}$ \\ The equation of the Wess-Zumino-Witten model is obtained by dropping the term in the right-hand-side of the first equation in (\ref{eq: pcm0}), \begin{equation} dz \wedge dA_z \wedge dp \wedge dq = 0 \ , \qquad d\bar z \wedge dA_{\bar z} \wedge dp \wedge dq = dz \wedge dA_z \wedge d\bar z \wedge dA_{\bar z} \ . \label{eq: wzw} \end{equation} After relabeling $(z,A_z,\bar z,A_{\bar z},p,q)$ as $(A_z,z,p,q,\bar z,A_{\bar z})$, (\ref{eq: wzw}) becomes \begin{equation} \partial_z A_{\bar z} - \partial_{\bar z} A_z= 0 \ , \qquad \{A_z, A_{\bar z} \} = 0 \ . \end{equation} Solving the potentials as in the case (b), we have \begin{equation} \Omega_{z, p} \Omega_{\bar z, q} - \Omega_{z, q} \Omega_{\bar z, p} = 0 \ . \end{equation} Comparing it with the first heavenly form (\ref{eq: pl}), we see that the WZW model describes a fully degenerate space-time \cite{park}. \vskip 0.2cm (d) ${\it The \ Higgs \ Bundle \ equation}$ \\ Here, note that it is not necessary to choose all four coordinate variables from the set $(z,x,\bar z,\bar x,p,q)$. Instead, we can regard more than two variables in it as functions. This observation enables us to obtain a further large class of models connected to the SdE. \par Let us examine the observation by taking $(z,\bar z)$ as two coordinate variables and $(x,\bar x,p,q) = (\phi,{\bar \phi},B_0,B_1)$ as functions. Then (\ref{eq: id}) reads \begin{equation} dz \wedge d\phi \wedge dB_0 \wedge dB_1 = dz \wedge d\phi \wedge d\bar z \wedge d{\bar \phi} \ , \quad d\bar z \wedge d{\bar \phi} \wedge dB_0 \wedge dB_1 = dz \wedge d\phi \wedge d\bar z \wedge d{\bar \phi} \ . \label{eq: hb} \end{equation} As like the case (b), we introduce $\partial_k A_{z(\bar z)} = \{B_0, B_1 \}_{(z(\bar z), x^k)}$, in which the coordinates $(p,q)$ are defined as the condition $\{B_0, B_1\}=1$ is satisfied. This condition ensures the integrability $\partial_{[k} \partial_{l]} A_{z(\bar z)} = 0$. Using $\phi, {\bar \phi}$ and $A_{z(\bar z)}$, (\ref{eq: hb}) and another identity ${\cal P} \wedge {\cal Q} \wedge {\cal P} \wedge {\cal Q} = 0$ become \begin{equation} \partial_{\bar z} \phi + \{A_{\bar z}, \phi \} = - \{\phi, \bar \phi \} \ , \quad \partial_{z} {\bar \phi} + \{A_z, {\bar \phi} \} = \{\phi, \bar \phi \} \ , \quad F_{z \bar z}(A) = 0 \ , \end{equation} or, changing $A_z \rightarrow A_z + \phi$ and $A_{\bar z} \rightarrow A_{\bar z} + {\bar \phi}$, \begin{equation} \partial_{\bar z} \phi + \{A_{\bar z}, \phi \} = 0 \ , \qquad \partial_{z} {\bar \phi} + \{A_z, {\bar \phi} \} = 0 \ , \qquad F_{z \bar z} = - \{\phi, {\bar \phi}\} \ . \end{equation} This is the two-dimensional Higgs bundle equation with the group SDiff(${\cal N}_2$), which is mentioned in Ref.\,\cite{ward}. A self-dual Einstein metric is given by a solution of the model through the tetrads, \begin{eqnarray} {\cal Z} = (\partial_p \phi + {g \over h} \partial_p \bar \phi) dz + {1 \over h} \partial_p \bar \phi \partial_{I} \phi \, dx^I \, , && {\cal \chi} = (\partial_q \phi + {g \over h} \partial_q \bar \phi) dz + {1 \over h} \partial_q \bar \phi \partial_{I} \phi \, dx^I \, , \\ {\bar {\cal Z}} = (\partial_q \bar \phi - {{\tilde g} \over h} \partial_q \phi) d\bar z - {1 \over h} \partial_q \phi \partial_{J} \bar \phi \, dx^J \, , && {\bar {\cal \chi}} = - (\partial_p \bar \phi - {{\tilde g} \over h} \partial_p \phi)d\bar z + {1 \over h} \partial_p \phi \partial_{J} \bar \phi \, dx^J \, , \nonumber \end{eqnarray} where $x^I$ and $x^J$ mean the sets of variables, $x^I=(\bar z,p,q)$, $x^J=(z,p,q)$, and $h= \{\phi,\bar \phi \}$, $g=\{\phi,A_z \}$ and ${\tilde g} = \{\bar \phi, A_{\bar z}\}$. The Higgs bundle equation was originally derived from a dimensional reduction of the four-dimensional self-dual Yang-Mills theory \cite{hitchin}. \vskip 0.2cm (e) Let us consider another example by choosing $(p,q)$ as two coordinates and $(z,x,\bar z,\bar x)= (C_0,C_1,D_0,D_1)$ as functions of $(p,q)$ and other $(z, \bar z)$. Also in this case we define $\partial_k A_{z(\bar z)} = \{C_0, C_1 \}_{(z(\bar z), x^k)}$, $\partial_k B_{z(\bar z)} = \{D_0, D_1 \}_{(z(\bar z), x^k)}$ and impose the conditions $\{C_0, C_1\}=1$, $\{D_0, D_1\}=1$ to ensure the integrability of $\partial_k A_{z(\bar z)}$ and $\partial_k B_{z(\bar z)}$. Then we obtain \begin{equation} F_{z \bar z}(A)= F_{z \bar z}(B)= 0 \ , \ \epsilon^{ab} (\partial_a B_b + \{A_a, B_b \} ) = 0 \ , \ \epsilon^{ab} (\partial_a A_b + \{B_a, A_b \} ) = 0 \ . \label{eq: new} \end{equation} The first two equations result from ${\cal Z} \wedge {\cal \chi} \wedge {\cal Z} \wedge {\cal \chi} = {\bar {\cal Z}} \wedge {\bar {\cal \chi}} \wedge {\bar {\cal Z}} \wedge {\bar {\cal \chi}}=0$. Tetrads in this case are, for example, \begin{eqnarray} &&{\cal Z} = h^{1 \over 2} dz - h^{- {1 \over 2}} \partial_k A_{\bar z} dx^k \ , \ \ \ {\cal \chi} = h^{1 \over 2}d\bar z + h^{- {1 \over 2}} \partial_k A_z dx^k \ , \ h=\{A_z,A_{\bar z}\} \ , \nonumber \\ &&{\bar {\cal Z}} = - g^{1 \over 2}d\bar z - g^{- {1 \over 2}} \partial_k B_z dx^k \ , \ {\bar {\cal \chi}} = g^{1 \over 2}dz - g^{- {1 \over 2}} \partial_k B_{\bar z} dx^k \ , \ \, g=\{B_z,B_{\bar z}\} \ . \end{eqnarray} In this model, the flat potentials $A_{z(\bar z)}$ and $B_{z(\bar z)}$ interact with each other through the third and fourth equations in (\ref{eq: new}). \vskip 0.4cm \section{Conclusion} % In this paper, we have shown a formulation of the self-dual Einstein space which leads to low-dimensional field theories quickly and clearly. Now the relation among those theories is rather clear. For example, if we obtain a solution of (a) the principal chiral model, then it is straightforward, at least formally, to derive the corresponding solution of (b) the topological model or the first heavenly form (\ref{eq: pl}); solving the potentials $A_{z(\bar z)}$ in (\ref{eq: pcm}) for $(p,q)$, relabeling $(z,A_z,\bar z,A_{\bar z},p,q)$ as $(z,\bar z,p,q,B_0,B_1)$ and following the step in the case (b). Here, for the opposite direction from (b) to (a), we give a simple example. A solution $\Omega$ of (\ref{eq: pl}) corresponding to the $k=0$ (flat) Gibbons-Hawking metric \cite{gibbons}\cite{EGH} is given by \begin{equation} \Omega = 2 \epsilon^{-1} \sqrt{\bar z q} \sinh{z} \sinh{p} + 2 \epsilon \sqrt{\bar z q} \cosh{z} \cosh{p} \ , \label{eq: ex} \end{equation} where $\epsilon$ is an arbitrary constant. We can make $\Omega$ real by setting the complex conjugate condition $z^{\ast}=p$, ${\bar z}^{\ast}=q$. Through the relation $\partial_k A_{z(\bar z)} = \{B_0, B_1 \}_{(z(\bar z), x^k)}$, the pair $(B_0, B_1)$ can take the form, \begin{equation} B_0 = \epsilon^{-1/2} \sqrt{2\bar z} \sinh{z} - \epsilon^{1/2} \sqrt{2q} \cosh{p} \ , \ B_1 = \epsilon^{1/2} \sqrt{2\bar z} \cosh{z} + \epsilon^{-1/2} \sqrt{2q} \sinh{p} \ . \label{eq: bb} \end{equation} By changing the notation as noted above and solving for $(A_z,A_{\bar z})$, we obtain a solution of the principal chiral model (\ref{eq: pcm}), \begin{eqnarray} A_z && = - [{{\epsilon^{-1/2} p \sinh{\bar z} + \epsilon^{1/2} q \cosh{\bar z} } \over {\epsilon^{-1} \sinh{z} \sinh{\bar z} + \epsilon \cosh{z} \cosh{\bar z}}}]^2 \ , \nonumber \\ A_{\bar z} && = - [{{\epsilon^{1/2} p \cosh{z} - \epsilon^{-1/2} q \sinh{z} } \over {\epsilon^{-1} \sinh{z} \sinh{\bar z} + \epsilon \cosh{z} \cosh{\bar z}}}]^2 \ . \end{eqnarray} Also it may be interesting to discuss various reduction procedures through this formulation. For example, in the case (b) with the above complex conjugate condition, suppose that $(A_z, A_{\bar z})$ are functions of $\bar z, q$ and the imaginary part of $(z,p)$ only. Then changing the imaginary part and $i A_z$, we obtain the three-dimensional Laplace equation, which is just the `translational' Killing vector reduction by Boyer and Finley \cite{boyer}. Also for the `rotational' case, the SL$(\infty)$ Toda equation is derived from (\ref{eq: id}) quickly. It is intriguing to pursue a further large class of models connected to the SdE by arranging the functions $(z,\bar z,x,\bar x,p,q)$ suitably and also to investigate the relation among the models. \par To check the coordinate variables which permit real, Euclidean metrices, let us write down the form of ${\cal P} \wedge {\cal Q}$ explicitly, \begin{equation} {\cal P} \wedge {\cal Q} = ({\cal Z} \wedge {\cal \chi} + {\bar {\cal Z}} \wedge {\bar {\cal \chi}}) + ({\cal Z} \wedge {\bar {\cal Z}} + {\cal \chi} \wedge {\bar {\cal \chi}}) \ . \label{eq: rc} \end{equation} If all tetrads are real, the real, Euclidean case, the first term in (\ref{eq: rc}) is real, while the second term pure-imaginary. In the case of (a) the principal chiral model, suppose $(p,q)$ are real or complex conjugate to each other. According to it, ${\cal P} \wedge {\cal Q} = dp \wedge dq $ becomes real or pure-imaginary. But then either the first or the second term in (\ref{eq: rc}) vanishes, which corresponds to a degenerate space-time, not an interesting case. Therefore, with such $(p,q)$, the chiral model permits only complex metric, or signature $(+,+,-,-)$ real metric in which case two of four tetrad one-forms may be taken as pure-imaginary. In fact $(B_0, B_1)$ in (\ref{eq: bb}), which are $(p,q)$ in the case (a), are neither real nor complex conjugate to each other. \par The infinite-dimensional group SDiff(${\cal N}_2$) is known to be realized as a large N limit of the SU(N) group when the surface ${\cal N}_2$ is the sphere or the torus \cite{sun}. Hence an approach to the SdE is to start from the SU(N) principal chiral model, there its explicit classical solutions can be determined by the Uhlenbeck's uniton construction \cite{uhlenbeck}. Adding to the SDiff(${\cal N}_2$), several (hidden) symmetrical structures in the SdE have been studied by the sigma model approach \cite{park}\cite{husain}\cite{morales}. Our formulation may be useful to investigate the structure since, from our key equations (\ref{eq: sm}),(\ref{eq: id}), we can easily see fundamental symmetries of our models, e.g. the conformal invariance on the $(z,\bar z)$ space and SDiff(${\cal N}_2$), before we derive their field equations explicitly. \newpage I am grateful to Q-Han Park, S. Nam, Ryu Sasaki and S. Odake for discussions. This work is supported by the department of research in Kyunghee University and the Japan Society for the Promotion of Science. \vskip 1.0cm
\section{Introduction} The recent discovery of anisotropy in the cosmic microwave background (CMB) lends support to the hypothesis that structure in the universe formed via gravitational instability from small density perturbations. Furthermore, the observed anisotropy allows us to characterize the initial perturbations statistically with enough precision to draw important cosmological conclusions. In particular, the COBE detection (Smoot et al.~1992, Bennett et al.~1994) shows that density perturbations existed on scales larger than the horizon at the epoch of recombination. Furthermore, the data are consistent with the hypothesis that the perturbations are Gaussian distributed and can be used to place constraints on the power spectrum of the initial density perturbations. CMB anisotropy is already one of the most important pieces of cosmological data, and in coming years its importance will only increase. In recent years, many cosmologists have favored theories in which the density parameter $\Omega_0$ is equal to one. In particular, this value for $\Omega_0$ is predicted by the simplest versions of the popular inflationary scenario. However, many dynamical measurements indicate a low-density universe (Peebles 1993, Ratra {$\&$ } Peebles 1994a). It is well known that the standard COBE-normalized Cold Dark Matter (CDM) scenario, in which universe is flat and contains only baryons and CDM particles, and in which the initial density perturbations are of the Harrison-Zel'dovich type, predicts fluctuations that are too large in amplitude on scales of galaxy clusters and below, although a number of slight variants on this model can be devised that fit the data better ({\it e.g.}, Bunn, Scott, \& White 1995; White et al.~1995). There are two classes of low-density cosmological model that can be motivated by the inflationary universe scenario. One is the $\Lambda$-model. In this model the matter density is small, so $\Omega_0<1$; however, the cosmological constant contributes to the total mean density required to make the universe flat: $\Omega_0+\Omega_\Lambda=1$. The usual inflationary scenarios predict that the universe is flat, so that $\Omega_0+\Omega_\Lambda$ must be extremely close to unity (Kashlinsky, Tkachev, {$\&$ } Frieman 1994). The cosmological constant is equal to the vacuum energy density of the Universe. Although we have no reason to be certain that it must be zero, the values that are of interest to cosmologists ({\it i.e.}, the values that make $\Omega_\Lambda$ of order unity) are unnaturally small from the point of view of particle physics (Weinberg 1989), and so these $\Lambda$-models are often regarded as unappealing. The other class of low-density inflationary scenario is the open model, in which the universe has negative spatial curvature. Recently, the possibility of realizing an open universe has been discussed in the context of inflation theory (Bucher, Goldhaber, {$\&$ } Turok 1994; Yamamoto, Sasaki, {$\&$ } Tanaka 1995; Linde 1995; Linde {$\&$ } Mezhlumian 1995). The essential idea is based on the semiclassical picture of a bubble nucleation, which is described by a bounce solution (Coleman {$\&$ } De~Luccia 1980). One bubble nucleation process can be regarded as the creation of a homogeneous and isotropic spacetime with negative spatial curvature inside the bubble due to the $O(4)$-symmetry of the bounce solution. It is of great interest to ask whether an open universe created in this one-bubble inflationary scenario is observationally acceptable or not. Open CDM models have been investigated by many authors (Lyth {$\&$ } Stewart 1990; Ratra {$\&$ } Peebles 1994ab; Sugiyama {$\&$ } Silk 1994; Kamionkowski et al. 1994; G\'orski et al. 1995; Liddle et al. 1995). Their investigations are based on the simple assumption that the quantum state of a scalar field is in the conformal vacuum state at the inflationary stage; however, this is unlikely to be the prediction of the one-bubble inflationary scenario. It has been pointed out that if we take the Bunch-Davies vacuum state as the state of scalar field, the CMB anisotropy in a low $\Omega_0$ universe appears quite different due to the super-curvature mode (Yamamoto, Sasaki, {$\&$ } Tanaka 1995). Following the usual inflationary picture in which the quantum fluctuation of a scalar field generates the density perturbation, we must investigate the quantum state of fields inside the bubble. Attempts have been made to study this problem by developing a field-theoretical formalism based on a multidimensional tunneling wave function (Tanaka, Sasaki, {$\&$ } Yamamoto 1994; Sasaki et al. 1994; Tanaka {$\&$ } Sasaki 1994; Yamamoto, Tanaka, {$\&$ } Sasaki 1995; Hamazaki et al. 1995). Bucher et al. have also considered this problem (Bucher, Goldhaber {$\&$ } Turok 1994; Bucher {$\&$ } Turok 1995). However, this problem requires further investigation. In this paper, we consider the simple case in which the quantum state of a scalar field is in the Bunch-Davies vacuum state, and compare the predictions of this model with several cosmological observations. This situation is physically definite and clear, and as long as the bubble nucleation occurs in the de Sitter inflationary background (where the initial inflationary period has lasted sufficiently long), the field can be approximated by the Bunch-Davies vacuum state, provided that the effect of the bubble nucleation process is negligible. In section 2, we first consider the initial spectrum by solving for the evolution of cosmological perturbations in the open inflationary stage. In section 3 we use this initial power spectrum to calculate various cosmological quantities and compare these predictions with observations. We also compare these results with those of a previous analysis of the open inflationary model based on the conformal vacuum state, and a $\Lambda$-model with Harrison-Zel'dovich spectrum. Section 4 is devoted to a discussion of our results. We will work in units where $c=1$ and $\hbar=1$. \section{Initial Conditions} In this section, we consider the evolution of cosmological perturbations in an open inflationary stage and derive the initial spectrum of perturbations. Ratra {$\&$ } Peebles (1994b) have investigated the evolution of cosmological perturbations in an open inflationary universe with gauge fixed. Bucher, Goldhaber {$\&$ } Turok (1994) have also investigated cosmological perturbations in an inflationary stage in the gauge-invariant formalism. In the first half of this section, we follow the work by Bucher, Goldhaber {$\&$ } Turok (1994). In an open universe the line element can be written as \begin{equation} ds^2=a^2(\eta)\biggl[ -d\eta^2+d\chi^2+\sinh^2\chi d\Omega^2_{(2)}\biggr], \label{metric} \end{equation} where $\eta$ is conformal time. We consider a scalar field $\phi$ with potential $V(\phi)$, which drives inflation. Writing the scalar field as a homogeneous part and a small inhomogeneous part, {i.e.}, $\phi=\phi_0+\delta\phi$, the equations for the homogeneous part are \begin{equation} \phi_0''+2{\cal H}\phi_0'+a^2 {\partial V\over\partial\phi}=0, \end{equation} and \begin{equation} {\cal H}^2 -1={8\pi G\over 3} \biggl[{\phi_0'^2\over 2}+a^2V\biggr], \end{equation} where ${\cal H}:=a'/a$ and the prime denotes differentiation with respect to $\eta$. The small fluctuation $\delta\phi$ gives rise to a metric perturbation. As we are interested in a scalar perturbation in a scalar field dominated universe, the metric perturbation can be written, \begin{equation} ds^2=a^2(\eta)\biggl[ -\bigl(1+2\Phi\bigr)d\eta^2+\bigl(1-2\Phi\bigr) \biggl(d\chi^2+\sinh^2\chi d\Omega^2_{(2)}\biggr) \biggr]. \end{equation} $\Phi$ corresponds to the curvature perturbation or gravitational potential. Then we get the evolution equations for the perturbation (Mukhanov, Feldman, {$\&$ } Brandenberger 1992), \begin{eqnarray} &&\Phi''+2\Bigl({\cal H}-{\phi_0''\over\phi_0'}\Bigr)\Phi' +\Bigl(-{\bf L}^2+4 +2{\cal H}'-2{\cal H}{\phi_0''\over\phi_0'}\Bigr)\Phi=0, \label{eqPhi} \\ &&\delta\phi={1\over4\pi G\phi_0'} \bigl(\Phi'+{\cal H}\Phi\bigr), \label{connect} \end{eqnarray} where \begin{equation} {\bf L}^2:={1\over\sinh^2\chi}{\partial\over\partial\chi} \biggl(\sinh^2\chi{\partial\over\partial\chi}\biggr)+ {1\over\sinh^2\chi}{\bf L}^2_\Omega, \end{equation} and ${\bf L}^2_\Omega$ is the Laplacian on the unit sphere. Now let us consider an inflationary stage of the universe inside a bubble. To solve the above equations analytically, we assume that the potential is nearly flat, and use the approximation, $V\simeq V_0+V'\phi$, where $V'={\rm const}$. We also assume that the background spacetime is approximated by de Sitter spacetime, that is, $a(t)=-1/H\sinh\eta$. Then the field equation for the homogeneous part is \begin{equation} \phi_0''-2{\rm coth}\,\eta\phi_0' = {-V' \over H^2\sinh^2\eta}, \label{phi0eq} \end{equation} with $H^2={8\pi G V_0/3}$. Eq. (\ref{phi0eq}) can be integrated, giving \begin{equation} \phi_0'={-V'\over H^2} {-\cosh^3\eta+3\cosh\eta\sinh^2\eta+2\sinh^3\eta\over 3\sinh\eta}, \end{equation} and the perturbation equation (\ref{eqPhi}) reduces to, \begin{equation} \Phi''-{6(1-e^{2\eta})\over 3-e^{2\eta}}\Phi'+ \biggl(-{\bf L}^2 +4-{4(3+e^{2\eta})\over3-e^{2\eta}}\biggr) \Phi=0. \label{eqPhib} \end{equation} To solve these equations, we need the initial values. To determine the initial values, we must investigate the problem of what the quantum state is inside the bubble. As mentioned before, this is a very important problem, which demands further investigation. We here consider the case in which the scalar field is in the Bunch-Davies vacuum state. This is the case provided that the effect of bubble nucleation process is small and negligible. But we should keep in mind that this point needs examination in the various models of one-bubble inflation scenario, taking into consideration the effect of bubble nucleation. Recently, quantum field theory in de Sitter space-time associated with the open chart has been investigated (Sasaki, Tanaka, {$\&$ } Yamamoto 1995). According to this analysis, a quantized scalar field with mass $m^2\ll H^2$ in the Bunch-Davies Vacuum state is described in the second quantized manner as, \begin{equation} \delta\phi=\int_0^\infty dp\sum_{\sigma,l,m} \chi_{p,\sigma}(\eta)Y_{plm}({\chi},\Omega_{(2)})\hat a_{p\sigma lm} + \sum_{lm}v_{(*)lm}(t,{\chi},\Omega_{(2)}) \hat a_{(*)lm} + {\rm h.c.}, \end{equation} where \begin{eqnarray} \chi_{\sigma,p}(\eta)&=&{-1\over\sqrt{8p(p^2+1)\sinh\pi p}} \nonumber \\ &&\biggl[e^{\pi p/2}(ip+\coth\eta)e^{-ip\eta} +\sigma e^{-\pi p/2}(ip-\coth\eta)e^{ip\eta} \biggr]{1\over a(\eta)}, \label{chi} \end{eqnarray} \begin{eqnarray} v_{(*)lm}(t,\chi,\Omega_{(2)}) &=&{H\over2}\sqrt{\Gamma(l+2)\Gamma(l)} {P^{-l-1/2}_{1/2}(\cosh\chi)\over \sqrt{\sinh\chi}} Y_{lm}(\Omega_{(2)}) \nonumber \\ &=:&{H\over2}W_{(*)l}(\chi)Y_{lm}(\Omega_{(2)}), {\hspace{14mm}} (l>0), \label{vstar} \end{eqnarray} $\sigma$ takes on the values $\pm1$, $\hat a$ is the annihilation operator, and $\Gamma(z)$ is the gamma function. The orthonormal harmonics on a three-dimensional unit hyperboloid $Y_{plm}(\chi,\Omega)$ are \begin{equation} Y_{plm}(\chi,\Omega_{(2)}) =\Bigg\vert{p\Gamma(ip+l+1)\over\Gamma(ip+1)}\Bigg\vert {P^{-l-1/2}_{ip-1/2}(\cosh\chi)\over \sqrt{\sinh\chi}} Y_{lm}(\Omega_{(2)}), \end{equation} with normalization \begin{equation} \int_0^\infty d\chi\int d\Omega_{(2)} \sinh^2\chi Y_{p_{1}l_{1}m_{1}}(\chi,\Omega_{(2)}) \overline{Y_{p_{2}l_{2}m_{2}}(\chi,\Omega_{(2)})} =\delta(p_{1}-p_{2}) \delta_{l_{1}l_{2}} \delta_{m_{1}m_{2}}. \end{equation} In the above expression, the usual harmonics behave as $Y_{plm}\propto e^{-\chi}$ at scales larger than the curvature scale, $\chi\gg1$, although $v_{(*)lm}$ is constant for $\chi\gg1$. Thus $v_{(*)lm}$ represents a fluctuation larger than the curvature scale, so we call this mode a super-curvature mode. The necessity of super-curvature modes for a complete description of a random field in an open universe has also been discussed by Lyth {$\&$ } Woszczyna (1995). Next, let us consider the curvature perturbation, which can be written in the mode-expanded form \begin{equation} \Phi=\int_0^\infty dp \sum_{\sigma,l,m} \Phi_{p,\sigma}(\eta) Y_{plm}(\chi,\Omega_{(2)})+ \sum_{l,m}\Phi_{(*)}(\eta)W_{(*)l}(\chi)Y_{lm}(\Omega_{(2)}). \end{equation} For the continuous mode $(p,l,m)$, Eq.(\ref{eqPhi}) reduces to \begin{equation} \Phi_p''-{6(1-e^{2\eta})\over 3-e^{2\eta}}\Phi_p'+ \biggl(p^2 +5-{4(3+e^{2\eta})\over3-e^{2\eta}}\biggr) {\Phi_p}=0. \end{equation} The solution that behaves like $\tilde\Phi_p\rightarrow e^{(1-ip)\eta}$ as $\eta\rightarrow -\infty$ is (Bucher {$\&$ } Turok 1995) \begin{equation} \tilde\Phi_p=e^{(1-ip)\eta}\biggl(1+{1+ip\over 1-ip} {e^{2\eta}\over3}\biggr). \end{equation} The equation for the super-curvature mode is \begin{equation} \Phi_{(*)}''-{6(1-e^{2\eta})\over 3-e^{2\eta}}\Phi_{(*)}'+ \biggl(4-{4(3+e^{2\eta})\over3-e^{2\eta}}\biggr) {\Phi_{(*)}}=0, \end{equation} and we find a solution \begin{equation} \tilde\Phi_{(*)}=e^{2\eta}. \end{equation} To determine the amplitude of $\Phi$, we use Eq.(\ref{connect}). From the behavior at $\eta\rightarrow-\infty$, we find for the continuous mode \begin{equation} \Phi_{p,\sigma}(\eta)={2\pi G V'\over H} {1\over\sqrt{8p(p^2+1)\sinh\pi p}} \biggl\{ {e^{\pi p/2}}{1-ip\over2-ip} \tilde\Phi_{p} +\sigma {e^{-\pi p/2}}{1+ip\over2+ip} \tilde\Phi_{-p}\biggr\}, \end{equation} and for the super-curvature mode \begin{equation} \Phi_{(*)}(\eta)={2\pi G V'\over H} {1\over3}\tilde\Phi_{(*)}(\eta). \end{equation} We therefore have the following spectrum at the end of inflation, by taking the limit $\eta\rightarrow0$, \begin{eqnarray} \Phi_p(0)^2 &:=&\lim_{\eta\rightarrow0} \sum_{\sigma=\pm1} \Phi_{p,\sigma}(\eta)^2 \nonumber \\ &=&\biggl({2\pi G V'\over H}\biggr)^2{\coth\pi p\over 2p(p^2+1)} {p^2+1\over p^2+4} \lim_{\eta\rightarrow0} \Bigl\vert \tilde\Phi_p(\eta)\Bigr\vert{}^2 \nonumber \\ &=&\biggl({4\pi G V'\over 3H}\biggr)^2 {\coth\pi p\over 2p(p^2+1)}, \label{specnm} \end{eqnarray} and \begin{eqnarray} \Phi_{(*)}(0)^2 &=&\lim_{\eta\rightarrow0}\Phi_{(*)}(\eta)^2 \nonumber \\ &=&\biggl({4\pi G V'\over 3H}\biggr)^2{1\over4}. \label{specsc} \end{eqnarray} For comparison, we also investigate the case when the scalar field is assumed to be in the conformal vacuum state (Lyth {$\&$ } Stewart 1990, Ratra {$\&$ } Peebles 1994b). In this case, the scalar field is written \begin{equation} \delta\phi=\int_0^\infty dp\sum_{l,m} \chi_{p}(\eta)Y_{plm}({\chi},\Omega_{(2)})\hat b_{plm} +{\rm h.c.} ~, \end{equation} where \begin{equation} \chi_{p}(\eta)={(ip+\coth\eta)e^{-ip\eta}\over\sqrt{2p(p^2+1)}} {1\over a(\eta)}. \end{equation} After a similar analysis, we get the spectrum of curvature perturbations at the end of inflation, \begin{equation} \lim_{\eta\rightarrow0}\Phi_p(\eta)^2 =\biggl({4\pi G V'\over 3H}\biggr)^2 {1 \over 2p(p^2+1)}. \label{spev} \end{equation} The conformal vacuum case differs from the Bunch-Davies vacuum case in two ways, the factor $\coth\pi p$ and the super-curvature mode. Lyth {$\&$ } Stewart (1990) have investigated perturbations in an open inflationary universe, and have given a relation to relate the curvature perturbation and the scalar field perturbation ${\cal R}\simeq -(H/\dot\phi)\delta\phi$, though a paper justifying this relation has never been published. But the above investigation shows the correctness of their result on all scales except for the small difference of the former coefficient. \section{Observational Confrontations} Now we start testing the predictions of the open universe in the context of CDM cosmology with the initial conditions obtained above. The matter-dominated open universe has the line element (\ref{metric}) with $a(\eta)=\cosh\eta-1$. In this section we use $\eta(>0)$ as the conformal time in the matter-dominated universe. \begin{center} \underline{(1) CMB Anisotropies} \end{center} Let us first consider the CMB temperature fluctuation. Having obtained an initial perturbation spectrum, we can compute the temperature fluctuations in the gauge-invariant formalism (Sugiyama {$\&$ } Gouda 1992). As usual, we write the temperature autocorrelation in the form, \begin{equation} C(\alpha)={1\over4\pi}\sum_{l}(2l+1)C_l{\rm P}_l(\cos\alpha). \end{equation} Figure 1(a) shows the power spectrum of temperature fluctuations, $l(l+1)C_l\times10^{10}/2\pi$, for various values of $\Omega_0$ with initial conditions associated with the Bunch-Davies vacuum state. We have taken $\Omega_B h^2=0.0125$ and Hubble parameter $h=0.75$, $0.70$, $0.65$, $0.65$, $0.60$, for $\Omega_0=0.1$, $0.2$, $0.3$, $0.4$, $0.5$, respectively, to take the age problem into consideration. Figure 1(b) shows the same quantities with the conformal vacuum state (Kamionkowski et al. 1994; G\'orski et al. 1995). For reference, we show the corresponding results for a $\Lambda$-model with a Harrison Zel'dovich spectrum in Figure 1(c) (Sugiyama 1995). Here the parameter $\Omega_B h^2=0.0128$ and $h=0.8$. We also show the results of several CMB experiments, taken from the paper by Scott, Silk {$\&$ } White (1995). Open models may have trouble fitting the data near the ``Doppler peak'' on degree scales, although assessing the significance of this problem will require very careful investigation (Ratra et al. 1995). The differences between Fig.1(a) and Fig.1(b) at low multipoles come almost entirely from the contribution of the super-curvature mode (Yamamoto, Sasaki, {$\&$ } Tanaka 1995), \begin{equation} C_{(*)l}=\biggl({2\pi G V'\over 3H}\biggr)^2 \biggl\{{1\over3}f(\eta_{LS})W_{(*)l}(\eta_0-\eta_{LS}) +2\int_{\eta_{LS}}^{\eta_0}d\eta' {df(\eta')\over d\eta' } W_{(*)l}(\eta_0-\eta') \biggr\}^2, \end{equation} where $\eta_{LS}$ and $\eta_0$ are the recombination time and the present time, respectively, $W_{(*)l}$ is defined in Eq.(\ref{vstar}), and $f(\eta)$ is the decay factor of the curvature perturbation, \begin{equation} f(\eta)=5~{\sinh^2\eta-3\eta\sinh\eta+4\cosh\eta-4\over(\cosh\eta-1)^3}. \label{feta} \end{equation} The most accurate and reliable CMB anisotropy data at the present time come from the COBE DMR experiment. In addition to providing us with accurate estimates of the fluctuation amplitude, the data from this experiment can be used to constrain the shape of the power spectrum. We have used the two-year COBE data (Bennett et al.~1994) to place constraints on open inflationary models, following a procedure based on the Karhunen-Lo\`eve transform (Bunn, Scott, \& White 1994; Bunn \& Sugiyama 1995; White \& Bunn 1995; Bunn 1995ab). This procedure gives results that are generally consistent with the spherical-harmonic technique devised by G\'orski (1994). We will now describe this procedure. In inflationary cosmological models, the CMB anisotropy is a realization of a Gaussian random field. If we expand the anisotropy in spherical harmonics, \begin{equation} \Delta T(\hat{\bf r}) = \sum_{l=2}^\infty\sum_{m=-l}^l a_{lm}Y_{lm} (\hat{\bf r}), \end{equation} then each coefficient $a_{lm}$ is an independent Gaussian random variable of zero mean. Furthermore, the variance of $a_{lm}$ is simply $C_l$. With this information, we can in principle compute the probability density $p(\vec d \,|\, C_l)$ of getting the actual COBE data $\vec d$ given a power spectrum $C_l$: since each data point is a linear combination of Gaussian random variables, the probability distribution $\vec d$ is simply a multivariate Gaussian, \begin{equation} p(\vec d \,|\, C_l)\propto\exp\left(-{1\over 2}\vec d^TM^{-1}\vec d\right), \label{GaussProb} \end{equation} where the covariance matrix $M$ can be written in terms of the power spectrum and the noise covariance matrix. Let us restrict our attention to a few-parameter family of possible power spectra. We will denote the parameters generically by $\vec q$. In this paper, for example, we will consider power spectra that are parameterized by two parameters, the density parameter $\Omega_0$ and the power spectrum normalization $Q\equiv\sqrt{5C_2/4\pi}$, and so $\vec q$ will be a two-dimensional vector. The probability density in equation (\ref{GaussProb}) is then simply the probability density $p(\vec d\,| \,\vec q)$ of the data $\vec d$ given the parameters $\vec q$. If we adopt a Bayesian view of statistics, we can convert this into a probability density for the parameters given the data: \begin{equation} p(\vec q \,|\, \vec d)\propto p(\vec d \,|\, \vec q) p(\vec q), \end{equation} where $p(\vec q)$ is the prior probability density we choose to adopt. $p(\vec q \,|\,\vec d)$ is generally denoted $L(\vec q)$ and called the likelihood. The choice of prior distribution is a notoriously troublesome issue. In practice, one generally chooses a prior that is a smooth, slowly-varying function of the parameters. In this paper, we will adopt a prior that is uniform in $\Omega_0$ and one that is uniform in $\ln Q$. (This prior is approximately equivalent to one that is uniform in the power spectrum normalization $C_{10}$ near the ``pivot point.'' It differs slightly from one that is uniform in $Q$, although not enough to affect our results significantly.) Unfortunately, in order to compute the probability density $p(\vec d\,|\,\vec q)$, and hence the likelihood $L$, one must invert a matrix of dimension equal to the number of data points. For the COBE DMR data, this number is of order 4000. Such exact likelihoods have been computed for a small class of models (Tegmark \& Bunn 1995); however, this is quite a time-consuming procedure. The Karhunen-Lo\`eve transform allows us to ``compress'' the data from 4000 numbers to only 400 in a way that throws away very little of the actual cosmological signal. The likelihoods estimated from the transformed data approximate the true likelihoods well, and are much more efficient to compute. For details on how the Karhunen-Lo\`eve transform is performed, see White \& Bunn (1995) and Bunn (1995ab). Once we know $L$, it is quite easy to place constraints on the parameters $\vec q$. Since $L$ is a probability distribution for $\vec q$, it should be normalized so that \begin{equation} \int L(\vec q)\,d\vec q = 1. \end{equation} Now suppose that we choose some subset $R$ of possible parameter values. Then if \begin{equation} \int_R L(\vec q)\,d\vec q=c \end{equation} then we can say that $\vec q$ lies in the region $R$ with probability $c$. If we want to find a 95\% confidence interval, we simply find a region $R$ such that $c=0.95$. One frequently chooses $R$ to be the region enclosed by a contour of constant likelihood. If one of the parameters is deemed to be uninteresting, the standard practice is to ``marginalize'' over it. For example, if we are interested in constraining $\Omega_0$ but not $Q$, then we replace $L(\Omega_0,Q)$ by \begin{equation} L_{\rm marg}(\Omega_0)=\int L(\Omega_0,Q)\,dQ. \end{equation} This is a natural thing to do: if $L$ is the joint probability density for $\Omega_0$ and $Q$, then $L_{\rm marg}$ is the probability density for $\Omega_0$ alone. Figure 2 shows the contours of the likelihood $L(\Omega_0,Q)$ for open models associated with the Bunch-Davies vacuum state. In computing these likelihoods, we use a linear combination of the 53 and 90 GHz two-year COBE maps, with weights chosen to minimize the noise. We use the ecliptic-projected maps; maps that were made in Galactic coordinates, and therefore have different pixelization, give normalizations that are generally lower by a few percent (Stompor, G\'orski \& Banday 1995; Bunn 1995a). The choice of pixelization appears to affect primarily the overall normalization of models; likelihood ratios of models with power spectra of different shapes are less affected (Bunn 1995ab). Figure 3 shows the marginal likelihoods for $\Omega_0$ for both the Bunch-Davies and the conformal vacuum open models. In the Bunch-Davies case, we find that $\Omega_0>0.34$ at 95\% confidence and $\Omega_0>0.15$ at 99\% confidence. We also show the confidence levels for various $\Omega_0$ in Table 1. For the conformal vacuum models, the likelihood is bimodal, and so the allowed regions are not connected. If we take a cut at small $\Omega_0$ and only consider the region $\Omega_0\geq0.03$, we can state that at 95\% confidence either $\Omega_0<0.085$ or $\Omega_0>0.36$, and 99\% confidence, either $\Omega_0<0.14$ or $\Omega_0>0.23$. There are difficulties associated with the interpretation of the likelihoods in this case (G\'orski et al. 1995). Of course, the likelihood $L(\Omega_0,Q)$ provides us with accurate normalizations in addition to shape constraints. For any particular value of $\Omega_0$, we find the value of $Q$ that maximizes the likelihood and use this value as the power spectrum normalization. The normalizations determined in this way have typical one-sigma fractional uncertainties of approximately $7.5\%$. The maximum-likelihood normalizations computed in this way are listed in the second column of Table 2(a) for the Bunch-Davies vacuum case. For comparison, we have also computed the conformal vacuum case; these normalizations are given in the second column of Table 2(b). \vspace{3mm} \begin{center} \underline{(2) Linear density power spectrum} \end{center} We next consider the matter inhomogeneities using the COBE normalization as described above. As the density perturbation $\Delta$ is related to the curvature perturbation $\Phi$ by the gravitational Poisson equation (Kodama {$\&$ } Sasaki 1984), \begin{equation} (p^2+4)\Phi_p(\eta)=4\pi G\rho(\eta) a^2(\eta)\Delta_p(\eta), \end{equation} in linear perturbation theory, we can write the power spectrum of the matter perturbation in an open universe from Eq.(\ref{specnm}), \begin{eqnarray} a_0^3P(k)~\Bigl(:=a_0^3\Delta_p^2\Bigr) &=&\biggl({2(1-\Omega_0)\over 3\Omega_0}\biggr)^2 (p^2+4)^2a_0^3\Phi_p^2(0) f^2(\eta_0) T(k)^2 \nonumber \\ &=:&{\cal A}(p^2+4)^2{\coth \pi p \over p(p^2+1)}T(k)^2, \label{Pk} \end{eqnarray} where $p=a_0 k$, $a_0=1/H_0\sqrt{1-\Omega_0}$, and $H_0=100h{\rm km/s/Mpc}$. The super-curvature mode does not contribute on small scales. The model based on the conformal vacuum state leads to the same form but without the factor $\coth\pi p$. In the CDM cosmology, the following transfer function is useful (Bardeen et al. 1986, Sugiyama 1995), \begin{equation} T(k)={\log(1+2.34q)\over 2.34q} \Bigl[1+3.89q+(16.1q)^2+(5.46q)^3+(6.71q)^4\Bigr]^{-1/4}, \label{transfersu} \end{equation} with \begin{equation} q=\Big({2.726\over2.7}\Bigr)^2 {k\over\Omega_0h\exp(-\Omega_B-\sqrt{2h}\Omega_B/\Omega_0)} ~h{\rm Mpc}^{-1} . \end{equation} This transfer function is for a flat universe; however, since we are interested in small-scale perturbations, the curvature of the universe can be neglected and the transfer function above is acceptable. The COBE DMR normalization determines the amplitude of the fluctuation. We give the numerical value of ${\cal A}$ determined from the likelihood normalization in the second column in Table 2(a). The second column in Table 2(b) gives the value for the conformal vacuum case, in which the power spectrum is obtained by Eq.(\ref{Pk}) without the factor $\coth\pi p$. We show in Figure 4(a) the density power spectrum $a_0^3P(k)$, for various $\Omega_0$, with initial conditions based on the Bunch-Davies vacuum state. The Hubble parameter is the same as that in Fig.1(a). The points are from Peacock {$\&$ } Dodds (1994). Figure 4(b) shows the density power spectra for $\Lambda$-models with Harrison-Zel'dovich power spectrum, in which the Hubble parameter is same as that in Fig.1(c). Given the density perturbation spectrum $P(k)$, we are able to calculate $\sigma_8^2=$$(\delta M$$/M)^2_{8 h^{-1}{\rm Mpc}}$, the variance of the mass fluctuation in a sphere of a radius $R=8h^{-1}{\rm Mpc}$, \begin{equation} \sigma^2(R)={1\over2\pi^2}\int k^2dk P(k) W^2(kR), \end{equation} where the top-hat window function $W$ is defined by $W(x)=3(\sin x -x\cos x)/x^3$. In Tables 2(a) and 2(b), we give $\sigma_8$ for various $\Omega_0$ and $ h$ in the open model associated with Bunch-Davies vacuum and the conformal vacuum respectively. We also show values for the $\Lambda$ model in Table 2(c) (Sugiyama 1995; Stompor et al. 1995). The difference between the Bunch-Davies case and the conformal vacuum case is very small. The difference is $10$ percent at $\Omega_0=0.05$, but is a few percent even at $\Omega_0=0.1$. This is because the COBE normalization based on a likelihood analysis gives more weight to the behavior of the power spectrum at $l\simeq10$, and less weight at lower multipoles (White {$\&$ } Bunn 1995). When we take a $\sigma(10^\circ)$ normalization, there is a $10$ percent difference between the two cases at $\Omega_0=0.1$. The values obtained are consistent with those in G\'orski et al. (1995). A precise comparison between predictions and observations of the matter power spectrum is difficult. One of the primary problems is that we do not know whether the galaxy distribution is an unbiased tracer of the mass distribution. However, if we make the reasonable assumptions that the galaxies are not anti-biased and are not extremely strongly biased (say $b\equiv\sigma_8^{-1}\mathrel{\mathpalette\oversim<} 2.5$), then these calculations suggest that $0.3\mathrel{\mathpalette\oversim<}\Omega_0\lsim0.5$. Note that the values of $\Omega_0$ preferred by the COBE likelihood analysis tend to be higher than this range; however, one might be inclined to argue that a model with $\Omega_0\simeq 0.4-0.5$ passes both tests. \vspace{3mm} \begin{center} \underline{(3)Large-scale bulk velocity} \end{center} Next, we consider large-scale bulk velocities, which are given by the following expression, \begin{equation} v_R^2= {H_0^2a_0^2\over 2\pi^2} \Omega_0^{1.2} \int dk P(k) W(kR)^2 \exp(-k^2R_s^2), \end{equation} where $W(kR)$ is the window function, and $R_s=12h^{-1}{\rm Mpc}$ is the Gaussian smoothing length for comparison with the observational data. In Table 3, we have summarized the computation of $v_R^2$ with $R=40h^{-1}{\rm Mpc}$ for various $\Omega_0$ for open models with the initial conditions associated with the Bunch-Davies vacuum state. These results are consistent with those of G\'orski et al. (1995). We can compare this results with the recent data from the POTENT analysis (Dekel~1994; Liddle et al.~1995): $v_{R=40h^{-1}{\rm Mpc}}=373\pm50{\rm km/s}$. Large values of $\Omega_0$ clearly provide a better fit to the velocities. It appears difficult to reconcile models with $\Omega_0\mathrel{\mathpalette\oversim<} 0.3$ with these data; however, it is difficult to make precise statistical statements based on these observations. As is discussed by Liddle et al. (1995), this measurement of the bulk velocity contains additional uncertainty due to cosmic variance. In addition, it is quite difficult to assess the uncertainties and potential biases in the POTENT analysis, and one should therefore be reluctant to draw firm conclusions on the basis of such a comparison. \vspace{3mm} \begin{center} \underline{(4) epoch of galaxy formation} \end{center} Liddle et al. (1995) have performed a detailed investigation of abundances of galaxy clusters and damped Lyman-alpha systems in open CDM models using Press-Schechter theory. In this paper we will rely on a simple and rough estimate of the epoch of galaxy formation, following the work of Gottl\"ober, M\"uchet, $\&$ Starobinsky (1994) and Peter, Polarski, $\&$ Starobinsky (1994). According to Press-Schechter theory, the fraction of the matter in the universe which is in gravitationally bound objects above a given mass $M_R$ at a redshift $z$ has the form \begin{equation} F(>M_R)={\rm erfc}\Bigl({\delta_c\over\sqrt{2}\sigma(M_R,z)}\Bigr), \end{equation} where \begin{equation} \sigma(M_R,z)=\sigma(R){1\over 1+z} {f(\eta_0)\over f(\eta(z))}, \end{equation} where $M_R=(4/3)\pi R^3\rho$, and $f(\eta)$ is the decay factor of the curvature perturbation. The choice of $\delta_c$ depends on the collapse model. The spherical collapse of a top-hat perturbation gives $\delta_c=1.69$, although non-spherical collapse models suggest other values. Here let us consider the range $(1.33<\delta_c<2)$ (Gottl\"ober et al. 1994). Observations suggest that many galaxies seems to have formed at $z=1$, then, assuming $F(>10^{12}{\rm M}_\odot) \mathrel{\mathpalette\oversim>} 0.1$ at $z=1$, we have $\sigma(M_R=10^{12}{\rm M}_\odot,z=1)\mathrel{\mathpalette\oversim>} 2\pm0.4$. Figure 5(a) shows a contour plot of $\sigma(M_R=10^{12}{\rm M}_\odot,z=1)$ in the $\Omega_0-h$ plane for the open model. Figure 5(b) is same but for the $\Lambda$-model with a Harrison Zel'dovich spectrum. If we take the age problem into consideration, it indicates a lower bound $\Omega_0\mathrel{\mathpalette\oversim>} 0.4$ for the open model. Note that the above estimate is very rough, although a similar constraint has been obtained from the exact estimation of cluster abundances (Liddle et al. 1995). The bound is weaker in the $\Lambda$-model than in the open model. \section{Discussion} A low-density universe is well motivated from several dynamical observations of galaxies and clusters. The simplest such low-density models are those in which the universe is open. In the context of inflation theory, however, we need a special idea such as the one-bubble inflationary scenario in order to produce an open universe. In this paper, motivated by the one-bubble inflationary universe scenario, we have examined the cosmological predictions based on the assumption that the scalar field is initially in the Bunch-Davies vacuum state. The initial perturbation spectrum has been derived by considering the evolution of perturbations in an open inflationary stage. Then the CMB anisotropies and the matter inhomogeneities have been examined. As the first test, we have performed a likelihood analysis for the CMB anisotropies by using the COBE DMR data. Interestingly, the COBE likelihood analysis gives severe constraints on the model. Models with $\Omega_0\leq0.4$, $\Omega_0\leq0.5$ are excluded at confidence levels of $92\%$, $83\%$, respectively. In a previous analysis associated with the conformal vacuum state (G\'orski et al. 1995), the likelihood function has another steep peak below $\Omega_0\lsim0.15$. This complicates the statistical interpretation of the results (G\'orski et al. 1995). In the case of the Bunch-Davies vacuum state, no such peak appears in the range of $\Omega_0$ we are interested in, and so the likelihood analysis gives clear results. The COBE likelihood analysis is therefore a powerful probe of these open models. We have used the the COBE DMR maximum-likelihood normalization to predict the amplitude of matter fluctuations. According to this normalization method, there is little difference between the predictions of the Bunch-Davies vacuum and conformal vacuum cases. Even for the case $\Omega_0=0.1$, the discrepancy of $\sigma_8$ is a few percent. We obtain results that are similar to previous analyses: the power spectrum of the mass fluctuation fits the observations of galaxies and clusters for $0.3\mathrel{\mathpalette\oversim<}\Omega_0\lsim0.5$. The required bias is unacceptably high for $\Omega_0\lsim0.1$, while high values of $\Omega_0$ demand anti-biasing. For example, $\Omega_0\gsim0.6$ needs anti-biasing when $h=0.65$. On the other hand, the $\Lambda$-models with Harrison Zel'dovich spectrum have higher amplitude compared with open models. The $\Lambda$-model needs anti-biasing for $\Omega_0\gsim0.4$ even when $h=0.65$. The $\Lambda$-model therefore needs low $h$ or a tilted spectrum (Ostriker {$\&$ } Steinhardt 1995). For the range $0.3\mathrel{\mathpalette\oversim<}\Omega_0\lsim0.5$, open models give unacceptably small bulk velocities compared with the POTENT analysis. However, given the present quality of the velocity data and the problem of cosmic variance, one might be reluctant to draw strong conclusions from this fact. The rough estimation associated with the galaxy formation gives lower bound of $\Omega_0$ consistent with the value discussed above. It is very interesting that the COBE likelihood analysis has given the most severe constraint on this open model. The COBE likelihood analysis strongly prefers a high value of $\Omega_0$. The peak value is around $0.7\mathrel{\mathpalette\oversim<}\Omega_0\lsim0.8$, and we can state that $\Omega_0\geq0.5$ with $83\%$ confidence in this model. Considering both the COBE analysis and the matter inhomogeneity, we are led to prefer a value of $\Omega_0\simeq 0.5$ if the one-bubble inflationary scenario is correct. Such a model is consistent with the Press-Schechter analysis of the epoch of galaxy formation and is marginally consistent with the bulk velocity data. As the CMB data continue to improve, particularly on degree scales, we should be able to test this model. It is premature to rule out low $\Omega_0$ inflationary models on the basis of this investigation at present, because we do not include the effect of bubble nucleation in the calculation of initial density power spectrum. Previous analysis indicates that the bubble nucleation effect in general excites fluctuations, and amplifies the perturbations on scales larger than curvature scale (Yamamoto, Tanaka, {$\&$ } Sasaki 1995; Hamazaki et al. 1995). One might therefore expect low-density models to fit the COBE data even more poorly once this effect is taken into account; however, since the calculation has not been done, we cannot be certain. In particular, the status of the super-curvature mode is still quite uncertain. Various modifications of the open model may also be viable. We must investigate the effect of gravity waves in an open inflationary universe. One might also consider the effect of tilting the primordial power spectrum; however, in order to improve the fit to the data one would probably need to tilt the power spectrum to have increased power on small scales, and such ``blue'' power spectra are not naturally produced by inflation. Such a model is probably too contrived to be plausible. \vspace{3mm} \begin{center} \bf Acknowledgments \end{center} We would like thank N. Sugiyama for providing us with the CMB anisotropy power spectra and for helpful discussions and comments. We are grateful to M. Sasaki and T. Tanaka for discussions and comments. One of us (K.Y.) would like to thank Professor J. Silk and the people at the Center for Particle Astrophysics, University of California, Berkeley, where many parts of this work were done for their hospitality. He would like to thank Professor H. Sato for continuous encouragement. This work was supported in part by Ministry of Education Grant-in-Aid for Scientific Research No. 2841. \vspace{10mm} \begin{center} \bf References \end{center} \noindent Bardeen, J. M., Bond, J. R., Kaiser, N., {$\&$ } Szalay, A. S. 1986, ApJ, {\bf 304}, 15. \\ Bennett, C.L, et al. 1994, ApJ, {\bf 436}, 423. \\ Bucher, M., Goldhaber, A. S., {$\&$ } Turok, N. 1994, PUTP-1507, hep-ph/9411206. \\ Bucher, M., {$\&$ } Turok, N. 1995, PUTP-1518, hep-ph/9503393. \\ Bunn, E.F. 1995a, Ph.D. dissertation, U.C. Berkeley Physics Department. \\ Bunn, E.F. 1995b, in preparation. \\ Bunn, E.F., Scott, D., \& White, M. 1995, ApJ, {\bf 441}, L9. \\ Bunn, E.F., \& Sugiyama, N. 1995, ApJ, in press. \\ Coleman, S., {$\&$ } De Luccia, F. 1980, Phys. Rev. {\bf D21}, 3305. \\ Dekel, A. 1994 ARA $\&$ A, 32, 371. \\ G\'orski, K. M. 1994, ApJ, {\bf 430}, L85. \\ G\'orski, K. M., Ratra, B., Sugiyama, N., {$\&$ } Banday, A. J. 1995, astro-ph/9502034, ApJ, in press. \\ Gottl\"ober, S., M\"uchet, J. P., {$\&$ } Starobinsky, A. A. 1994, ApJ, {\bf 434}, 417. \\ Hamazaki, T., Sasaki, M., Tanaka, T., {$\&$ } Yamamoto, K. 1995, KUNS1340. \\ Kamionkowski, M., Ratra, B., Spergel, D. N., {$\&$ } Sugiyama, N. 1994, ApJ, {\bf 434}, L1. \\ Kashlinsky, A., Tkachev, A. A., {$\&$ } Frieman, J. 1994, Phys. Rev. Lett., {\bf 73}, 1582. \\ Kodama, H., {$\&$ } Sasaki, M. 1984, Prog. Theor. Phys. Suppl. {\bf 78}, 1. \\ Liddle, A. R., Lyth, D. H., Roberts, D., {$\&$ } Voana, P. T. 1995, SUSSEX-AST 95/6-2, astro-ph/9506091, MNRAS, in press. \\ Linde, A. 1995, preprint SU-ITP-95-5, hep-th/9503097, Phys. Lett. {\bf B}, in press. \\ Linde, A., {$\&$ } Mezhlumian, A. 1995, preprint SU-ITP-95-11, astro-ph/9506017. \\ Lyth, D. H., {$\&$ } Stewart, E. D. 1990, Phys. Lett. {\bf B252}, 336. \\ Lyth, D. H., {$\&$ } Woszczyna, A. 1995, Lancaster preprint, astro-ph/9408069. \\ Mukhanov, V. F., Feldman, H. A., {$\&$ } Brandenberger, R. H. 1992, Phys. Rep. {\bf 215}, 203. \\ Ostriker, J. P., {$\&$ } Steinhardt, P. J. 1995, astro-ph/9505066. \\ Peacock, J. A., {$\&$ } Dodds, S. J. 1994, MNRAS, 267, 1020. \\ Peebles, P. J. E. 1993, {\it Principle of Physical Cosmology}, Princeton Univ. Press. \\ Peter, P., Polarski, D., {$\&$ } Starobinsky, A. A. 1994, Phys. Rev. {\bf D50}, 4827. \\ Ratra, B., et al. 1995, in preparation. \\ Ratra, B., {$\&$ } Peebles, P. J. E. 1994a, ApJ, {\bf 432}, L5. \\ Ratra, B., {$\&$ } Peebles, P. J. E. 1994b, preprint PUPT-1444. \\ Sasaki, M., Tanaka, T., Yamamoto, K., {$\&$ } Yokoyama, J. 1994, Prog. Theor. Phys. 90, 1019. \\ Sasaki, M., Tanaka, T., {$\&$ } Yamamoto, K. 1995, Phys. Rev. {\bf D51}, 2979. \\ Scott, D., Silk, J., {$\&$ } White, M. 1995, preprint. \\ Smoot, G. F. et al. 1995, ApJ. {\bf 396}, L1. \\ Stompor, R., G\'orski, K. M., {$\&$ } Banday, A. J. 1995, preprint. \\ Sugiyama, N. 1995, preprint CfPA-TH-94-62. \\ Sugiyama, N., {$\&$ } Gouda, N. 1992, Prog. Theor. Phys., 88, 803. \\ Sugiyama, N., {$\&$ } Silk, J. 1994, Phys. Rev. Lett., {\bf 73}, 509. \\ Tanaka, T., Sasaki, M., {$\&$ } Yamamoto, K. 1994, Phys. Rev. {\bf D49}, 1039. \\ Tanaka, T., {$\&$ } Sasaki, M. 1994, Phys. Rev. {\bf D50}, 6444. \\ Tegmark, M., \& Bunn, E. F. 1995, Berkeley preprint. \\ Weinberg, S. 1989, Rev. Mod. Phys., 61, 1. \\ White, M., {$\&$ } Bunn, E. F. 1995, CfPA-95-TH-02, ApJ, in press. \\ White, M., Scott, D., Silk, J., \& Davis, M. 1995, preprint. \\ Yamamoto, K., Sasaki, M., {$\&$ } Tanaka, T. 1995, KUNS-1309, ApJ, in press. \\ Yamamoto, K., Tanaka, T., {$\&$ } Sasaki, M. 1995, Phys. Rev. {\bf D51}, 2968. \def{Q}{{Q}} \def{\mu {\rm K}}{{\mu {\rm K}}} \vspace{12mm} \centerline{Table 1} \begin{centerline} {Confidence levels for Bunch-Davies open model } \end{centerline} \begin{center} \begin{tabular}{c@{\hspace{1pc}}c} \hline\hline {\ \ $\Omega_0(>)$ \ \ }&{confidence level $(\%)$} \\ \hline $ 0.1$ & $99.4$ \\ $ 0.2$ & $98.4$ \\ $ 0.3$ & $96.2$ \\ $ 0.4$ & $91.8$ \\ $ 0.5$ & $83.2$ \\ \hline \end{tabular} \end{center} \vspace{2mm} \centerline{Table 2(a)} \begin{centerline} {Amplitude of density perturbation for Bunch-Davies open model } \end{centerline} \begin{center} \begin{tabular}{c@{\hspace{1pc}}c@{\hspace{1pc}}c@{\hspace{1pc}}c@{\hspace{1pc}}c} \hline\hline {\ \ $\Omega_0$ \ \ }&{$\cal A$} & {\ \ } & {$\sigma_8$} & {\ \ } \\ {\ \ } & {$(h^{-1}{\rm Mpc})^3$} & {h=0.5} & {h=0.65} & { h=0.8} \\ \hline $ 0.05$ & $1.96({Q}/27.9{\mu {\rm K}})^2~\times10^{2}$ & $0.0041$ & $0.011$& $0.021$ \\ $ 0.1 $ & $2.42({Q}/28.8{\mu {\rm K}})^2~\times10^{2}$ & $0.032$ & $0.063$ & $0.099$ \\ $ 0.2 $ & $2.83({Q}/27.8{\mu {\rm K}})^2~\times10^{2}$ & $0.15$ & $0.25$ & $0.35$ \\ $ 0.3 $ & $2.94({Q}/25.8{\mu {\rm K}})^2~\times10^{2}$ & $0.31$ & $0.48$ & $0.66$ \\ $ 0.4 $ & $2.90({Q}/23.4{\mu {\rm K}})^2~\times10^{2}$ & $0.49$ & $0.74$ & $1.00$ \\ $ 0.5 $ & $2.72({Q}/21.1{\mu {\rm K}})^2~\times10^{2}$ & $0.69$ & $1.01 $ & $1.33$ \\ $ 0.6 $ & $2.43({Q}/19.3{\mu {\rm K}})^2~\times10^{2}$ & $0.89$ & $1.26 $ & $1.63$ \\ $ 0.7 $ & $2.03({Q}/18.3{\mu {\rm K}})^2~\times10^{2}$ & $1.05$ & $1.45 $ & $1.90$ \\ $ 0.8 $ & $1.54({Q}/18.3{\mu {\rm K}})^2~\times10^{2}$ & $1.20$ & $1.66 $ & $2.10$ \\ $ 0.9 $ & $1.00({Q}/18.9{\mu {\rm K}})^2~\times10^{2}$ & $1.31$ & $1.80 $ & $2.25$ \\ \hline \end{tabular} \end{center} \newpage \centerline{Table 2(b)} \begin{centerline} {Amplitude of density perturbation for conformal vacuum open model } \end{centerline} \begin{center} \begin{tabular}{c@{\hspace{1pc}}c@{\hspace{1pc}}c@{\hspace{1pc}}c@{\hspace{1pc}}c} \hline\hline {\ \ $\Omega_0$ \ \ } & {$\cal A$} & {\ \ } & {$\sigma_8$} & {\ \ } \\ {\ \ } & {$(h^{-1}{\rm Mpc})^3$} & {h=0.5} & {h=0.65} & { h=0.8} \\ \hline $ 0.05$ & $2.45({Q}/18.7{\mu {\rm K}})^2~\times10^{2}$ & $0.0045$ &$0.012$ & $0.023$ \\ $ 0.1 $ & $2.54({Q}/23.1{\mu {\rm K}})^2~\times10^{2}$ & $0.032$ & $0.064$ & $0.10$ \\ $ 0.2 $ & $2.89({Q}/26.5{\mu {\rm K}})^2~\times10^{2}$ & $0.15$ & $0.25$ & $0.36$ \\ $ 0.3 $ & $3.01({Q}/25.9{\mu {\rm K}})^2~\times10^{2}$ & $0.31$ & $0.49$ & $0.67$ \\ $ 0.4 $ & $2.96({Q}/23.5{\mu {\rm K}})^2~\times10^{2}$ & $0.50$ & $0.75$ & $1.01$ \\ $ 0.5 $ & $2.77({Q}/20.6{\mu {\rm K}})^2~\times10^{2}$ & $0.69$ & $1.02$ & $1.34$ \\ $ 0.6 $ & $2.46({Q}/18.3{\mu {\rm K}})^2~\times10^{2}$ & $0.88$ & $1.27$ & $1.64$ \\ $ 0.7 $ & $2.04({Q}/17.2{\mu {\rm K}})^2~\times10^{2}$ & $1.06$ & $1.50$ & $1.90$ \\ $ 0.8 $ & $1.55({Q}/17.5{\mu {\rm K}})^2~\times10^{2}$ & $1.20$ & $1.67$ & $2.10$ \\ $ 0.9 $ & $1.00({Q}/18.7{\mu {\rm K}})^2~\times10^{2}$ & $1.31$ & $1.80$ & $2.25$ \\ \hline \end{tabular} \end{center} \vspace{5mm} \centerline{Table 2(c)} \begin{centerline} {Amplitude of density perturbation for Harrison-Zel'dovich $\Lambda$-model } \end{centerline} \begin{center} \begin{tabular}{c@{\hspace{1pc}}c@{\hspace{1pc}}c@{\hspace{1pc}}c} \hline\hline {\ \ $\Omega_0$ \ \ } & {\ \ } & {$\sigma_8$} & {\ \ } \\ {\ \ } & {h=0.5} & {h=0.65} & { h=0.8} \\ \hline $ 0.1 $ & $0.15$ & $0.29$ & $0.46$ \\ $ 0.2 $ & $0.41$ & $0.68$ & $0.98$ \\ $ 0.3 $ & $0.65$ & $1.0$ & $1.4$ \\ $ 0.4 $ & $0.85$ & $1.3$ & $1.7$ \\ $ 0.6 $ & $1.2 $ & $1.7$ & $2.2$ \\ \hline \end{tabular} \end{center} \vspace{5mm} \centerline{Table 3} \begin{centerline} {Large scale bulk velocity for Bunch-Davies open model } \end{centerline} \begin{center} \begin{tabular}{c@{\hspace{1pc}}c@{\hspace{1pc}}c@{\hspace{1pc}}c} \hline\hline {\ \ $\Omega_0$ \ \ } & {\ \ } & {$v_{R=40h^{-1}{\rm Mpc}} ({\rm km/s})$ } & {\ \ } \\ {\ \ \ \ } & {h=0.5} & {h=0.65} & { h=0.8} \\ \hline $ 0.05$ & $6.3$ & $8.6$ & $ 11$ \\ $ 0.1 $ & $ 21$ & $ 29$ & $ 37$ \\ $ 0.2 $ & $ 71$ & $ 90$ & $106$ \\ $ 0.3 $ & $130$ & $160$ & $180$ \\ $ 0.4 $ & $200$ & $230$ & $260$ \\ $ 0.5 $ & $260$ & $300$ & $330$ \\ $ 0.6 $ & $320$ & $360$ & $390$ \\ $ 0.7 $ & $370$ & $410$ & $450$ \\ $ 0.8 $ & $410$ & $450$ & $490$ \\ $ 0.9 $ & $440$ & $480$ & $510$ \\ \hline \end{tabular} \end{center} \newpage \begin{center} \underline{Figure Captions} \end{center} \noindent {Figure 1.} Power spectra of the CMB temperature anisotropy $l(l+1)C_l\times10^{10}/2\pi$ for (a) the Bunch-Davies vacuum open model with $\Omega_0=0.1,0.2,0.3,0.4,0.5$. The data were provided by N. Sugiyama. These theoretical curves are normalized by the COBE likelihood. The curves are in descending order of $\Omega_0$ as one moves down at $l=50$. The results of several CMB experiments are also shown, taken from the paper by Scott, Silk, {$\&$ } White (1995). To compare with degree-scale observations, careful investigations are required (Ratra et al. 1995). \vspace{3mm} \noindent {Figure 1(b).} CMB power spectra for the conformal vacuum open model (G\'orski et al. 1995). The curves and points are as in Figure 1(a). \vspace{3mm} \noindent {Figure 1(c).} Power spectrum of the CMB temperature anisotropy $l(l+1)C_l\times10^{10}/2\pi$ for the Harrison-Zel'dovich $\Lambda$-model with $\Omega_0=0.1,0.2,0.3,0.4,1.0$. (Sugiyama 1995) \vspace{3mm} \noindent {Figure 2.} Contour plot of the likelihood function $L(\Omega_0,Q)$ for the Bunch-Davies open model. The contour range is from $L=0.25$ to $L=1.5$, where the likelihoods are scaled so that $L=1$ corresponds to a flat Harrison-Zel'dovich spectrum with maximum-likelihood normalization. \vspace{3mm} \noindent {Figure 3.} The marginal likelihood $L_{\rm marg}(\Omega_0)$ as a function of $\Omega_0$ for both the Bunch-Davies open model (solid line) and the conformal vacuum open model (dashed line). \vspace{3mm} \noindent {Figure 4(a).} Power spectrum of density perturbation $a_0^3 P(k)$ for the open model with $\Omega_0=0.1$, $0.2$, $0.3$, $0.4$, $0.5$, for $h=0.75, 0.70, 0.65, 0.65, 0.60$, respectively. We have taken $\Omega_Bh^2=0.0125$. The points are from Peacock {$\&$ } Dodds (1994). \vspace{3mm} \noindent {Figure 4(b).} Same figure as Fig.4(a) but for the Harrison-Zel'dovich $\Lambda$-model with $\Omega_0=$ $0.1$, $0.2$, $0.3$, $0.4$, $1.0$. We have taken $\Omega_Bh^2=$$0.0128$ and $h=0.8$. \vspace{3mm} \noindent {Figure 5(a).} Contours of $\sigma(M_R=10^{12} {\rm M}_\odot,z=1)$, in the $(\Omega_0-h)$ plane for the Bunch-Davies open models. The contour range is from $\sigma=1.0$ to $\sigma=6.0$. The dashed lines are $\sigma=1.6$ and $2.4$. \vspace{3mm} \noindent {Figure 5(b).} Same figure as Fig.5(a) but for Harrison-Zel'dovich $\Lambda$-models. \end{document}
\section*{\large\bf 1. Introduction} In molecular dynamics simulations of heavy-ion collisions at intermediate energies (300 MeV/u -- 2 GeV/u) it is still an open question question, which kind of forces should be used for the long range part of the nucleon-nucleon interaction which builds up the mean field. The Skyrme forces, which work very well for lower energies are not relativistically invariant. There are attempts to extend non-relativistic two-body potentials to higher velocities by requiring approximate Lorentz invariance~\cite{Fel1,Schm89,Stac76,Bodm80}. Our aim is to derive from a relativistic lagrangian of the Walecka type a covariant Lagrange function for point-like particles which is suited for molecular dynamics and still has the saturation property of the original lagrangian. It should then be applicable in transport theories as for example QMD or BUU and should give reasonable results even at lower energies (50 -- 200 MeV/u). In Section~2 the relativistic Hamilton function is derived in the small acceleration approximation \cite{Weber} from a many-body Lagrange function and it is expressed in terms of positions and canonical momenta of the particles. In Section~3 the QMD procedure is described, while in Section~4 we present our QMD calculations for fragment distributions in O + Br collisions at different energies and compare the calculated results with the experimental data. A few years ago the analogue calculation was done~\cite{kn:judqmd} with the Skyrme interaction, which is explicitly density dependent. It is interesting to compare those results with the new forces presented in this paper in order to see how far they can describe saturation and low energy phenomena. The intention is, however, to go to intermediate energies, where QMD is more appropriate because quantum aspects like the Pauli principle or the uncertainty relation are expected to be less important. \section*{\large\bf 2. Relativistic equations of motion for nucleons} In this section we are deriving the relativistic equations of motion for the mean 4-positions and the mean 4-momenta of the nucleons. We start from a field theoretical lagrangian of the following type \begin{eqnarray}\label{Lagr1} {\cal L}(x) & = & \bar{\psi}(x) \Big( \gamma^\alpha i\partial_\alpha - m^*\left( \phi(x)\right) \Big) \psi(x) \nonumber \\ & - & g_{\rm v} \bar{\psi}(x) \gamma^\alpha \psi (x) A_\alpha (x) \nonumber \\ & - & \frac{1}{2} \phi(x) (\partial_\alpha \partial^\alpha + \mu^2_{\rm s}) \phi (x) \nonumber \\ & + & \frac{1}{2} A^\alpha(x) (\partial_\beta \partial^\beta + \mu^2_{\rm v}) A_\alpha (x) \ , \end{eqnarray} where we allow for two different ways to couple the scalar field~\cite{Lindner}. If $m^*(\phi) = m - g_{\rm s} \phi$ we obtain the Walecka lagrangian, whereas for $m^*(\phi) = \left( m + g_{\rm s} \phi \right)^{-1}$ we deal with the lagrangian proposed by Zim\'anyi and Moszkowski~\cite{kn:zimmos}. (In the following, if not expressed explicitly otherwise, we use units such that $\ \hbar = c = 1$.) The field equations are \begin{equation}\label{Dirac} \Big\{(i \partial_\alpha - g_{\rm v} A_\alpha(x)) - m^*\left( \phi(x)\right)\Big\}\ \psi(x) = 0 \end{equation} for the nucleons, \begin{equation}\label{Wavephi} (\partial^\beta \partial_\beta + \mu^2_{\rm s})\ \phi(x) = -\ {d \over{ d \phi }} m^* \left( \phi(x) \right) \ \bar{\psi}(x)\psi(x) \end{equation} for the scalar field and \begin{equation}\label{WaveA} (\partial^\beta \partial_\beta + \mu_{\rm v}^2)\ A^\alpha(x) = g_{\rm v} \, \bar{\psi}(x) \gamma^\alpha \psi(x) \end{equation} for the vector field. In the mean-field approximation the source terms are replaced by their expectation values $\bar{\psi}(x)\psi(x) \Rightarrow \rho_{\rm s}(x)$ and $\bar{\psi}(x) \gamma^\alpha \psi(x) \Rightarrow j^\alpha(x)$ so that the scalar and vector fields become classical fields. For classical fields, however, an arbitrarily weak time dependence in the scalar density $\rho_{\rm s}(x)$ or current density $j^\alpha(x)$ results in a radiation field which travels away from the nucleons. The total field energy in the radiation may even be less than the masses of the field quanta. Since the fields are only effective fields, and for the scalar field there is no corresponding elementary particle, the physical meaning of this radiation is obscure. As we are using the relativistic scalar and vector fields only for the mean-field part of the nuclear interaction we follow the suggestion \cite{Fel1,Fel2} to exclude radiation of fictitious mesons from the very beginning by using the action-at-a-distance formulation with the symmetric Green's function of Wheeler and Feynman \cite{Wheeler}. \begin{equation} G_{\rm s,v}(x-y) = \frac{1}{2} \Big( G^{advanced}_{\rm s,v} (x-y)+ G^{retarded}_{\rm s,v} (x-y)\Big) \ \end{equation} The formal solution of the wave Eq.~(\ref{Wavephi}) (for $m^* = m - g_{\rm s} \phi$) \begin{equation}\label{Phigreens} \phi(x) = g_{\rm s} \int\! d^4 y \ G_{\rm s} (x-y) \bar{\psi}(y)\psi(y) \end{equation} and of Eq. (\ref{WaveA}) \begin{equation}\label{Agreens} A^\alpha(x) = g_{\rm v} \int\! d^4 y \ G_{\rm v} (x-y) \bar{\psi}(y)\gamma^\alpha\psi(y) \end{equation} fulfills the desired boundary condition of vanishing incoming and outgoing free fields. Eliminating the fields from the lagrangian (\ref{Lagr1}) leads to the non-local action which contains only nucleon variables: \begin{eqnarray}\label{Action} \int d^4x\ {\cal L}(x) & = & \int d^4x\ \bar{\psi}(x) (\gamma^\alpha i\partial_\alpha - M)\psi(x) \nonumber \\ &+& \frac{1}{2}g_{\rm s}^2 \int d^4x\ d^4y\ \bar{\psi}(x)\psi(x)\, G_{\rm s}(x-y)\, \bar{\psi}(y)\psi(y) \nonumber \\ &-& \frac{1}{2}g_{\rm v}^2 \int d^4x\ d^4y\ \bar{\psi}(x)\gamma^\alpha\psi(x) \,G_{\rm v}(x-y)\, \bar{\psi}(y)\gamma_\alpha\psi(y) \end{eqnarray} If quantum effects are negligible the scalar density and vector current density can be represented in terms of the world lines of the nucleons as \begin{eqnarray}\label{rhos} \rho_{\rm s}(x):= \bar{\psi}(x)\psi(x) &\rightarrow& \sum_{\rm j=1}^{\rm A} \int\! d \tau^{\ }_{\rm j} \ \delta^4 (x-r_{\rm j}(\tau^{\ }_{\rm j})) \\ j^\mu(x):= \bar{\psi}(x)\gamma^\mu\psi(x) &\rightarrow& \sum_{\rm j=1}^{\rm A} \int\! d \tau^{\ }_{\rm j} \ \delta^4 (x-r_{\rm j}(\tau^{\ }_{\rm j}))\, u^\mu_{\rm j}(\tau^{\ }_{\rm j})\ , \label{rhos2} \end{eqnarray} where $r_{\rm j}(\tau^{\ }_{\rm j})$ denotes the world line of the nucleon $j$ at its proper time $\tau^{\ }_{\rm j}$ and $u_{\rm j}(\tau^{\ }_{\rm j})$ its 4-velocity. Nucleons in nuclei can however not be localized well enough for treating them as classical particles on world lines. The Fermi momentum sets a limit to the radius of the nucleon wave packet in coordinate space which is of the order of 2 fm. These effects, the Pauli principle and other quantum effects, are properly taken care of in Antisymmetrized Molecular Dynamics (AMD)~\cite{kn:amd} and in Fermionic Molecular Dynamics (FMD)~\cite{kn:fmd}. However, presently neither the AMD nor the FMD equations of motion can be solved numerically for large systems like gold on gold. Therefore, we mimic the finite size effect of the wave packets by folding the finally obtained forces with gaussian density distributions for the nucleons. This "smearing" provides also the prescription how to eliminate strong forces at short distances which are taken care of by the random forces of the collision term. Therefore, the forces to be derived in the following are only meant to represent the long range part of the nuclear interactions which constitute a mean field. In this sense $r_{\rm j}(\tau^{\ }_{\rm j})$ is the mean world line of the nucleon and $u_{\rm j}(\tau^{\ }_{\rm j})$ its mean 4-velocity. Using the representation (\ref{rhos}) the scalar field at a space-time point $x$, for example is given by integrals over past and future proper times $\tau^{\ }_{\rm j}$ as \begin{equation}\label{Phiworld} \phi(x) = g_{\rm s} \sum_{\rm j=1}^{\rm A} \int\! d\tau^{\ }_{\rm j} \ G_{\rm s}(x-r_{\rm j}(\tau^{\ }_{\rm j})) \ . \end{equation} By identifying the following expressions with their representation in terms of mean positions and mean momenta of the nucleons \begin{eqnarray} \int d^4x\ d^4y\ \bar{\psi}(x)\psi(x)\,G_{\rm s}(x-y)\, \bar{\psi}(y)\psi(y) &\rightarrow& \nonumber \\ && \mbox{\hspace*{-6cm}} \sum^{\rm A }_{\stackrel{{\scriptstyle \rm i,j=1}}{\rm i\neq j}} \int\! d\tau_i d\tau^{\ }_{\rm j}\ G_{\rm s}(r_{\rm i}(\tau^{\ }_{\rm i})- r_{\rm j}(\tau^{\ }_{\rm j})) \quad ,\\ \int d^4x\ d^4y\ \bar{\psi}(x)\gamma^\alpha\psi(x) \,G_{\rm v}(x-y)\, \bar{\psi}(y)\gamma_\alpha\psi(y) &\rightarrow& \nonumber \\ &&\mbox{\hspace*{-6cm}} \sum^{\rm A}_{\stackrel{{\scriptstyle \rm i,j=1}}{\rm i\neq j}} \int\! d\tau^{\ }_{\rm i} d\tau^{\ }_{\rm j}\ G_{\rm v}(r_{\rm i}(\tau^{\ }_{\rm i})- r_{\rm j}(\tau^{\ }_{\rm j}))\, u_{{\rm i}\alpha}(\tau^{\ }_{\rm i}) u_{\rm j}^\alpha(\tau^{\ }_{\rm j}) \end{eqnarray} in the action (\ref{Action}) one obtains the non-instantaneous action-at-a-distance \cite{Wheeler} \begin{eqnarray}\label{Action1} \int\! d^4x \ {\cal L}(x) &\rightarrow& \nonumber \\ {\cal A} &=& - \frac{1}{2} \sum^{\rm A}_{\rm i=1} \int\! d\tau^{\ }_{\rm i} \left(M-\lambda_{\rm i}(\tau^{\ }_{\rm i})\right) u_{\rm i}(\tau^{\ }_{\rm i})^2 \nonumber \\ &+&\frac{1}{2} g^2_{\rm s} \sum^{\rm A}_{\stackrel{{\scriptstyle \rm i,j=1}}{\rm i\neq j}} \int\! d\tau^{\ }_{\rm i} d\tau^{\ }_{\rm j} \ G_{\rm s}(r_{\rm i}(\tau^{\ }_{\rm i}) - r_{\rm j}(\tau^{\ }_{\rm j})) \\ &-&\frac{1}{2} g^{2}_{\rm v} \sum^{\rm A}_{\stackrel{{\scriptstyle \rm i,j=1}}{\rm i\neq j}} \int\! d\tau^{\ }_{\rm i} d\tau^{\ }_{\rm j} \ G_{\rm v}(r_{\rm i}(\tau^{\ }_{\rm i}) - r_{\rm j}(\tau^{\ }_{\rm j}))\ u_{{\rm i}\alpha}(\tau^{\ }_{\rm i}) u_{\rm j}^\alpha(\tau^{\ }_{\rm j}) \ .\nonumber \end{eqnarray} Lagrange multipliers $\lambda_{\rm i}(\tau^{\ }_{\rm i})$ are introduced to ensure $u_{\rm i}(\tau^{\ }_{\rm i})^2=1$. The equations of motion which result from this Wheeler-Feynman action are equivalent to the classical field equations, provided the system does not radiate \cite{Wheeler}. However, the Wheeler Feynman equations of motion cannot be solved in general as one needs to know the world lines for all the past and future times in order to calculate the fields which enter the equations for the world lines~\cite{BelMartin}. Furthermore, the no-interaction theorem states that, except in the case of non-interacting particles, there exist no covariant equations of motion for world lines in which only the 4-positions and the 4-velocities at a given time enter, as is the case in non-relativistic mechanics. In the following this no-interaction theorem is circumvented by introducing an approximative solution to the non-local Wheeler Feynman equations of motion. This is achieved by the so called small acceleration approximation which does not assume small velocities. \subsection*{\large\sl 2.1. Action-at-a-distance in the small acceleration approximation} In order to conserve manifest covariance of the equations of motion we introduce first the concept of a scalar time. For that the whole Minkowski space is chronologically ordered by a set of space like surfaces $S(x)$ which attribute to each 4-position $x$ a scalar time $t$ by \begin{equation}\label{Isochrone} S(x)=t \ \ , \ \ \ \mbox{with} \ \ \ \partial^0 S(x) > 0 \ ,\ \partial^0 \partial^0 S(x) \ge 0 \ . \end{equation} The simplest choice for these isochrones, which we shall take in the following, is $S(x)=\eta_\alpha x^a$, where $\eta_\alpha$ is a position independent time like 4-vector. Unlike in a collision the nucleons are not strongly accelerated by the action of the mean-field. Therefore, in order to describe the motion in the mean-field one may expand each world line around the proper time $\tau^s_{\rm j}$ at which the particle is at a given isochrone, i.e. $t=S(r_{\rm j}(\tau^s_{\rm j}))$ for each $j$. \begin{equation}\label{worldline} r_{\rm j}^\alpha(\tau^{\ }_{\rm j})=r_{\rm j}^\alpha(\tau^s_{\rm j}) +(\tau^{\ }_{\rm j}-\tau^s_{\rm j}) u_{\rm j}^\alpha(\tau^s_{\rm j})+ \frac{1}{2}(\tau^{\ }_{\rm j}-\tau^s_{\rm j})^2 \, a_{\rm j}^\alpha(\tau^s_{\rm j})+\cdots \ . \end{equation} This way we are defining a synchronization prescription for all particles, which does not depend on the frame. For calculating the fields we neglect of the quadratic term with the acceleration $a_{\rm j}^\alpha(\tau^s_{\rm j})$ and all higher powers which leaves us with a straight world line in the vicinity of the synchronizing time $t=S(r_{\rm j}(\tau^s_{\rm j}))$ (c.f. Fig.~1). This is called ''small acceleration approximation" \begin{figure}[t] \vspace*{-10mm} \insertplot{ff-wline.ps} \vspace*{-10mm} \caption{World lines and synchronizing hypersurface $S(x)$} \end{figure} The small acceleration approximation is of course best in the vicinity of $r_{\rm j}^\alpha(\tau^s_{\rm j})$ and becomes worse further away. As sketched in Fig. 1 only those parts of the world lines (thick lines), which are outside the light cone (centered at $x$), contribute to the field strength at point $x$. A world line which hits the light cone far away from $x$ may be badly approximated by Eq. (\ref{worldline}), but for short range interactions a distant particle does not contribute anymore to the field at $x$. Thus, the first condition for the validity of the approximation is that the range $\mu^{-1}$ is small compared to the curvature of the world lines, i.e. the inverse of the acceleration. The second condition is weak radiation, which is fulfilled when the acceleration is small compared to the meson mass $\mu$. Both conditions are actually the same, namely \begin{equation}\label{smallacc} \mid a_\mu a^\mu \mid \ \ll \ \mu^2 . \end{equation} The assumption of small accelerations is justified if the $\phi$ and $A^\alpha$ fields are only meant to be the mean-field part of the nucleon-nucleon interaction in a hadronic surrounding. The hard collisions between individual nucleons which are due to the repulsive core will cause large accelerations and also create new particles. These hard collisions cannot be described within Walecka type mean-field models. Therefore, it is consistent to regard $\phi (x)$ and $A^\alpha(x)$ as Hartree mean-fields which bring about only small accelerations and which are not radiated away from their sources. Inserting the straight world line into Eq. (\ref{Phiworld}) results in the easily understood situation that the field at a point $x$ is just the sum of Lorentz-boosted Yukawa potentials which are traveling along with the charges: \begin{equation}\label{Phiyuk} \phi(x) = \frac{g_{\rm s}}{4\pi} \sum_{\rm j} \frac{\exp\left\{-\mu_{\rm s} \sqrt{R_{\rm j}(x)^2}\right\} } {\sqrt{R_{\rm j}(x)^2}}\ , \end{equation} where $R_{\rm j}(x)^2$ is given by \begin{equation} R_{\rm j}(x)^2 = - (x-r_{\rm j}(\tau^s_{\rm j}))^2 + \left[(x^\alpha-r_{\rm j}^\alpha(\tau^s_{\rm j}))\, u_{{\rm j}\alpha}(\tau^s_{\rm j})\right]^2 \end{equation} The vector field is derived in an analogue fashion as \begin{equation}\label{Ayuk} A^\alpha(x) = \frac{g_{\rm v}}{4\pi} \sum_{\rm j} \frac{\exp\left\{-\mu_{\rm v}\sqrt{R_{\rm j}(x)^2}\right\} } {\sqrt{R_{\rm j}(x)^2}}\, u_{\rm j}^\alpha(\tau^s_{\rm j}) \ . \end{equation} At this level of the approximation the causality problem with the advanced part of the Green function is not present because the retarded and the advanced fields are identically the same when they are created by charges moving on straight world lines. Therefore one may regard the fields as retarded only. In addition the unsolved problem of radiation reaction, where the radiation acts back on the world lines \cite{Jackson}, does not occur because there are no radiation fields any more. \subsection*{\large\sl 2.2. Instantaneous action-at-a-distance} In the spirit of the small acceleration assumption discussed in the previous section one can use the straight line expansion in the action (\ref{Action1}) and perform the integration over $\tau^{\ }_{\rm j}$. This results in an instantaneous action-at-a-distance where the Lorentz-boosted Yukawa fields appear again and there is only one time, the scalar synchronizing time $t$. \begin{eqnarray}\label{Action2} {\cal A} = \int \! dt \sum_{\rm i} \frac{1}{\partial_\alpha S(r_{\rm i}(t))u_{\rm i}^\alpha(t)} \Bigg[ &-&\frac{1}{2} \left(M-\lambda_{\rm i}(t)\right)u_{\rm i}(t)^2 \nonumber\\ &+& \frac{1}{2} \sum_{\stackrel{\scriptstyle\rm j}{\rm j \neq i}} \frac{g^2_{\rm s}}{4\pi} \frac{\exp\{-\mu_{\rm s} \sqrt{R_{\rm ij}(t)^2}\}}{\sqrt{R_{\rm ij}(t)^2}} \\ && \mbox{\hspace*{-2cm}}- \frac{1}{2} \sum_{\stackrel{\scriptstyle\rm j}{\rm j \neq i}} \frac{g^2_{\rm v}}{4\pi} \frac{\exp\{-\mu_{\rm v} \sqrt{R_{\rm ij}(t)^2}\}}{\sqrt{R_{\rm ij}(t)^2}}\, u_{{\rm i}\alpha}(t) u_{\rm j}^\alpha(t)\,\Bigg] \nonumber \end{eqnarray} The four-positions $r_{\rm i}^\alpha(t)\equiv r_{\rm i}^\alpha(\tau^{\ }_{\rm i}(t))$ and 4-velocities $u_{\rm i}^\alpha(t) \equiv u_{\rm i}^\alpha(\tau^{\ }_{\rm i}(t))$ are to be taken at the same scalar synchronizing time $t$ and \begin{equation} R_{\rm ij}^2(t) := - (r_{\rm i}(t) - r_{\rm j}(t))^2 + \Big[(r_{{\rm i}\beta}(t) - r_{{\rm j}\beta}(t)) u_{\rm j}^\beta (t)\Big]^2 \end{equation} The small acceleration approximation together with the introduction of a synchronizing hypersurface $S(x)$ leads to an equal time lagrangian which is Lorentz-scalar and written in a manifestly covariant way. Giving up explicit covariance and choosing $\eta=(1,0,0,0)$ in a special coordinate frame, the positions and velocities take the form \begin{equation}\label{eq:rcm} r_{\rm i}(\tau^{\ }_{\rm i}(t)) = \Big(t\, , \,\vec{r}_{\rm i}(t)\Big) \ \ \ \mbox{and} \ \ \ u_{\rm i}(\tau^{\ }_{\rm i}(t)) = \frac{1}{\sqrt{1-\vec{v}_{\rm i}^{\,2}(t)}} \Big( 1\, ,\, \vec{v}_{\rm i}(t) \Big) \ . \end{equation} With that a Lagrange function ${\cal L} (\vec{r}_{\rm i}(t), \vec{v}_{\rm i}(t))$ can be defined which depends only on the independent variables and one time. Even the Lagrange multipliers $\lambda_{\rm i}(t)$ are not needed anymore if the variation is with respect to $\vec{v}_{\rm i}(t)$ instead of all four $u_{\rm i}^\alpha(t)$. The advantage of the instantaneous Lagrangian is that one can define easily the hamiltonian and the total momentum, which are then strictly conserved by the equations of motion. \subsection*{\large\sl 2.3. Hamilton equations of motion} In the following we want to express the total hamiltonian as a function of the positions $\vec{r}_{\rm i}$ and the canonical momenta $\vec{p}_{\rm i}$. Using relations (\ref{Action2})-(\ref{eq:rcm}) the Lagrange function is \begin{eqnarray} \label{eq:lang} {\cal L} = &-&\sum_{\rm i} \ \left[ {m \over{ u_{\rm i}^0 }} + {1 \over 2} \sum_{\stackrel{\scriptstyle\rm j}{\rm j \neq i}}\ {g^2_{\rm s} \over{ 4\pi u_{\rm i}^0 }} {{\rm exp}\left\{ -\mu_{\rm s} \sqrt{R_{\rm ij}^2(t)} \right\} \over{ \sqrt{R_{\rm ij}^2(t)} }} \right. \\ &-& \left. {1 \over 2} \sum_{\stackrel{\scriptstyle\rm j}{\rm j \neq i}}\ {g^2_{\rm v} \over{ 4\pi u_{\rm i}^0 }} {{\rm exp}\left\{ -\mu_{\rm v} \sqrt{R_{\rm ij}^2(t)} \right\} \over{ \sqrt{R_{\rm ij}^2(t)} }} u_{\rm i\alpha}(t) u_{\rm j}^{\alpha}(t) \right] \, . \nonumber \end{eqnarray} The canonical momenta and the energy are determined from the Lagrange function above as \begin{eqnarray} \vec{p}_{\rm i} &=& {\partial {\cal L} \over{ \partial \vec{v}_{\rm i} }} \quad , \\ E &=& \sum_{\rm i} {\partial {\cal L} \over{ \partial \vec{v}_{\rm i} }} \vec{v}_{\rm i} - \cal{L} \quad .\nonumber \end{eqnarray} Performing these derivatives, the energy is given by \begin{eqnarray} \label{eq:enu} E &=& \sum _{\rm i} \ m\ u_{\rm i}^0 - {1\over 2} \sum_{\stackrel{\scriptstyle\rm i,j}{\rm j \neq i}}\ f_{\rm ij} u_{\rm i}^0 - {1\over 2} \sum_{\stackrel{\scriptstyle\rm i,j}{\rm j \neq i}}\ \widehat{f}_{\rm ji}\ (\vec{r}_{\rm ij} \vec{u}_{\rm i})^2 {\left( u_{\rm i}^0 \right)^2 \over{ u_{\rm j}^0 }} \nonumber \\ &+& {1\over 2} \sum_{\stackrel{\scriptstyle\rm i,j}{\rm j \neq i}}\ g_{\rm ji}\ u_{\rm i}^0 - {1\over 2} \sum_{\stackrel{\scriptstyle\rm i,j}{\rm j \neq i}}\ g_{\rm ji} \left[ u_{\rm i}^0 \vec{u}_{\rm i}^2 - {\left( u_{\rm i}^0 \right)^2 \over{ u_{\rm j}^0 }} \vec{u}_{\rm i} \vec{u}_{\rm j} \right] \\ &+& {1\over 2} \sum_{\stackrel{\scriptstyle\rm i,j}{\rm j \neq i}}\ \widehat{g}_{\rm ji}\ (\vec{r}_{\rm ij} \vec{u}_{\rm i})^2 \left[ \left( u_{\rm i}^0 \right)^3 - {\left( u_{\rm i}^0 \right)^2 \over{ u_{\rm j}^0 }} \vec{u}_{\rm i} \vec{u}_{\rm j} \right] , \nonumber \end{eqnarray} \noindent where we use the abbreviations \begin{eqnarray} \label{eq:fdef1} f_{\rm ij} \equiv f(R_{\rm ij}) = {g_{\rm s}^2\over{4 \pi}}\ {{\rm exp}\left\{ -\mu_{\rm s} \rm R_{ij}\right\} \over{R_{\rm ij}}} \quad &,& \quad g_{\rm ij} = {g_{\rm v}^2\over{4 \pi}}\ {{\rm exp}\left\{ -\mu_{\rm v} \rm R_{ij}\right\} \over{R_{\rm ij}}} \, , \nonumber \\ \widehat{f}_{\rm ji} \equiv -{1 \over{ R_{\rm ji} }} {d \over{ d R_{\rm ji} }} f(R_{\rm ji}) = {g_{\rm s}^2\over{4 \pi}}\ (1 &\hspace*{-5mm}+& \hspace*{-5mm} \mu_{\rm s} R_{\rm ji}) {{\rm exp}\left\{ -\mu_{\rm s} \rm R_{ji}\right\} \over{ R_{\rm ji}^3 }}\, , \\ R_{\rm ij}^2 = \vec{r}_{\rm ij}^2 + (\vec{r}_{\rm ij} \vec{u}_{\rm j})^2 \quad &,& \quad \vec{r}_{\rm ij} = \vec{r}_{\rm i} - \vec{r}_{\rm j} \, , \label{eq:fdef2} \end{eqnarray} and $\widehat{g}_{\rm ji}$ is defined in the same way as $\widehat{f}_{\rm ji}$. It is worthwhile to note that $f_{\rm ij} \neq f_{\rm ji}$. The momentum of a particle looks as \begin{eqnarray} \label{eq:mom} \vec{p}_{\rm i} &=& m\vec{u}_{\rm i} - {1\over 2} \left( \sum_{\stackrel{\scriptstyle\rm j}{\rm j \neq i}}\ f_{\rm ij} \right) \vec{u}_{\rm i} - {1\over 2} \sum_{\stackrel{\scriptstyle\rm j}{\rm j \neq i}}\ \widehat{f}_{\rm ji} {u_{\rm i}^0 \over{ u_{\rm j}^0 }} \left[ \vec{r}_{\rm ij} (\vec{r}_{\rm ij} \vec{u}_{\rm i}) + \vec{u}_{\rm i} (\vec{r}_{\rm ij} \vec{u}_{\rm i})^2 \right] \nonumber \\ &+& {1\over 2} \sum_{\stackrel{\scriptstyle\rm j}{\rm j \neq i}}\ \left( g_{\rm ij} + {u_{\rm i}^0 \over{ u_{\rm j}^0 }} g_{\rm ji} \right) \vec{u}_{\rm j} - {1\over 2} \sum_{\stackrel{\scriptstyle\rm j}{\rm j \neq i}}\ g_{\rm ji} \left[ \left( u_{\rm i}^0 \right)^2 - {u_{\rm i}^0 \over{ u_{\rm j}^0 }} \vec{u}_{\rm i} \vec{u}_{\rm j} \right] \vec{u}_{\rm i} \\ &+& {1\over 2} \sum_{\stackrel{\scriptstyle\rm j}{\rm j \neq i}}\ \widehat{g}_{\rm ji} \left[ \left( u_{\rm i}^0 \right)^2 - {u_{\rm i}^0 \over{ u_{\rm j}^0 }} \vec{u}_{\rm i} \vec{u}_{\rm j} \right] \left[ \vec{r}_{\rm ij} (\vec{r}_{\rm ij} \vec{u}_{\rm i}) + \vec{u}_{\rm i} (\vec{r}_{\rm ij} \vec{u}_{\rm i})^2 \right] \, . \nonumber \end{eqnarray} Let us first prove that for isotropic nuclear matter, like in the original mean-field picture~\cite{kn:serwal} the vector potential will not contribute to the canonical momentum. In that case the summation can be replaced by an integration over space and momentum, folded with the phase space distribution function. Since for isotropic nuclear matter the distribution function is independent of the position, the space integral can be carried out immediately and the remaining part is written as a summation over momenta. This way we get for the Yukawa terms in nuclear matter \begin{eqnarray} \sum_{\rm j}\ f_{\rm ij} &\rightarrow& \sum_{\rm j}\ \int d^3r_{\rm j} \ f(R_{\rm ij}) = \left( {g_{\rm s} \over{ \mu_{\rm s}}} \right)^2 \sum_{\rm j}\ {1\over{u_{\rm j}^0}} \nonumber \, . \\ \sum_{\rm j}\ \widehat{f}_{\rm ji}\ (\vec{r}_{\rm ij} \vec{u}_{\rm i})^2 &\rightarrow& \sum_{\rm j}\ \int d^3r_{\rm j}\ \widehat{f}(R_{\rm ji}) (\vec{r}_{\rm ij} \vec{u}_{\rm i})^2 = \left( {g_{\rm s} \over{ \mu_{\rm s}}} \right)^2 {\vec{u}_{\rm i}^2\over{\left( u_{\rm i}^0 \right)^3 }} \ \sum_{\rm j}\ 1 \, . \label{eq:nms} \end{eqnarray} Substituting the above expressions into the momentum~(\ref{eq:mom}) the $\widehat{f}_{\rm ij}$ term gives the same contribution as the $f_{\rm ij}$, while the $g_{\rm ij}$ terms cancel each other and the momentum can be written as \begin{equation} \vec{p}_{\rm i} = m \vec{u}_{\rm i} - \left( {g_{\rm s}\over{ \mu_{\rm s} }} \right)^2 \vec{u}_{\rm i} \sum_{\stackrel{{\scriptstyle\rm j}}{\rm j\neq i}} \ {1\over{u_{\rm j}^0}} + \left( {g_{\rm v}\over{\mu_{\rm v} }} \right)^2 \sum_{\stackrel{{\scriptstyle\rm j}}{\rm j\neq i}} \ {\vec{u}_{\rm j}\over{ u_{\rm j}^0}} \, . \label{eq:pnm0} \end{equation} The last term gives zero, since the average of $\vec{u}_{\rm j}$ is zero in nuclear matter and Eq.~(\ref{eq:pnm0}) can be written as \begin{equation} \vec{p}_{\rm i} = m^* \vec{u}_{\rm i} \quad , \label{eq:pnm1} \end{equation} \noindent where \begin{eqnarray} m^* &=& m - \left( {g_{\rm s}\over{\mu_{\rm s} }} \right)^2 \sum_{\stackrel{{\scriptstyle\rm j}}{\rm j\neq i}} \ {1\over{u_{\rm j}^0}} \nonumber \\ &=& m - \left( {g_{\rm s}\over{\mu_{\rm s} }} \right)^2 \sum_{\rm j} \ {m^* \over{\sqrt{{m^*}^2 + \vec{p}_{\rm j}^2}}} \, . \label{eq:msnm} \end{eqnarray} Similarly, the energy density (\ref{eq:enu}) has the form \begin{equation} \epsilon = \sum_{\rm i}\ \sqrt{{m^*}^2 + \vec{p}_{\rm i}^2} + {1\over 2} \left( {g_{\rm v}\over{\mu_{\rm v} }} \right)^2 \left( j^0 \right)^2 + {1\over 2} \left( {\mu_{\rm s} \over{ g_{\rm s} }} \right)^2 (m - m^*)^2 \quad . \label{eq:ednm} \end{equation} \noindent Eqs.~(\ref{eq:msnm}-\ref{eq:ednm}) are just the mean-field results of the Walecka model~\cite{kn:serwal}. It is worthwhile to mention that the small acceleration approximation which introduces $R_{\rm ij}$ instead of $\mid \vec{r}_{\rm ij}\mid$ in Eqs.~(\ref{eq:fdef1}-\ref{eq:fdef2}), is definitely needed to get back the relativistic mean-field result for nuclear matter. Therefore, for the coupling strengths $g_{\rm s}$ and $g_{\rm v}$ we can simply take the values obtained from the saturation properties of the nuclear matter~\cite{Lindner}. For the Hamilton equations we need the energy as the function of the canonical momenta instead of the 4-velocities $u_{\rm i}$. For that we have to invert the $p(u)$ equation (\ref{eq:mom}) into a $u(p)$ one. One can do this approximatively by first expanding the 4-velocities in the $p(u)$ relation up to the third order in the quantities $(\vec{p}/m)^2$, $f_{\rm ij}$ and $g_{\rm ij}$, and then substituting that expression for $u(p)$ into the energy Eq.~(\ref{eq:enu}). At the end we will check the validity of this approximation. After tedious calculations the energy can be expressed as follows (in the following we include explicitly $c$ and $\hbar$) \begin{eqnarray} \label{eq:en} {E\over{mc^2}} &=& \sum_{\rm i}\ \sqrt{1+\vec{\tilde{p}}_{\rm i}^2} - {1\over 2} \sum_{\stackrel{{\scriptstyle \rm i,j}}{\rm j \neq i}}\ {\tilde{f}_{\rm ij} \over{ \sqrt{1+\vec{\tilde{p}}_{\rm i}^2}}} + {1\over 2} \sum_{\stackrel{{\scriptstyle \rm i,j}}{\rm j \neq i}}\ \tilde{g}_{\rm ji} \sqrt{1+\vec{\tilde{p}}_{\rm i}^2} \nonumber \\ &-& {1\over 2} \sum_{\stackrel{{\scriptstyle \rm i,j}}{\rm j \neq i}}\ \tilde{g}_{\rm ji} {(\vec{\tilde{p}}_{\rm i} \vec{\tilde{p}}_{\rm j}) \over{ \sqrt{1+\vec{\tilde{p}}_{\rm i}^2}}} + {1\over 2} \sum_{\rm i}\ \left( \sum_{\stackrel{{\scriptstyle \rm j}}{\rm j \neq i}}\ \tilde{f}_{\rm ji} \vec{\tilde{p}}_{\rm i} - \sum_{\stackrel{{\scriptstyle \rm j}}{\rm j \neq i}}\ \tilde{g}_{\rm ji} \vec{\tilde{p}}_{\rm j} \right)^2 \quad , \end{eqnarray} \noindent where \begin{eqnarray} \tilde{f}_{\rm ij} &=& {1\over{4\pi}} \left( {g_{\rm s} m \over{ \mu_{\rm s}}} \right)^2 \left( {\hbar \over{ mc }} \right) {{\rm exp}\left\{ -{\hbar \over{ \mu_{\rm s} c }} \tilde{R}_{\rm ij}\right\} \over{\tilde{R}_{\rm ij}}} \quad {\rm and} \nonumber \\ \tilde{R}_{\rm ij}^2 &=& \vec{r}_{\rm ij}^2 + ( \vec{r}_{\rm ij} \vec{\tilde{p}}_{\rm j} )^2 \quad , \end{eqnarray} \noindent with $\vec{\tilde{p}}_{\rm i} = \vec{p}_{\rm i}/mc$. $\tilde{g}_{\rm ij}$ is defined similarly to $\tilde{f}_{\rm ij}$. The last term is already of third order and hence small. {}From Eq.~(\ref{eq:en}) the Hamilton equations \begin{equation} {d \over{ dt }} \vec{r}_{\rm i} = -\ {\partial E\over{\partial \vec{p}_{\rm i}}} \qquad {d \over{ dt }} \vec{p}_{\rm i} = {\partial E\over{\partial \vec{r}_{\rm i}}} \end{equation} are calculated easily. \section*{\large\bf 3. Details of the QMD calculations} Molecular dynamics is a classical many-body theory in which some quantum features due to the fermionic nature of nucleons are simulated. For the details of the theory we would like to refer the reader to the works of Aichelin~\cite{kn:aich} and the Frankfurt group~\cite{kn:frank}. In our calculation we introduced some modifications which are described in~\cite{kn:judqmd,kn:paula}. In the following this model is used in a relativistic treatment with scalar and vector forces. \subsection*{\large\sl 3.1. The mean-field forces} The scalar and vector forces are the relativistic two-body forces derived in the previous section which give saturation for nuclear matter even without an explicit density dependent term. However, Walecka's values for the coupling constants result in a too high compressibility. For this reason it seems to be more adequate to use instead the Zim\'anyi-Moszkowski (ZM) lagrangian~\cite{kn:zimmos} which provides a more reasonable compressibility. The different coupling of the mean-field ($m^* = ( m + g_{\rm s} \phi )^{-1}$) modifies the scalar part of the total energy. In the used approximation, the expression \begin{equation} -{1\over 2} \sum_{\stackrel{\scriptstyle \rm i,j}{ \scriptstyle \rm j \neq i}} \ {\tilde{f}_{\rm ij} \over{ \sqrt{1 + \vec{\tilde{p}}_{\rm i}^2} }} \left( {1 \over{ 1 + 2 {\displaystyle \sum_{\stackrel{\scriptstyle \rm k}{\rm k \neq j}}\ \tilde{f}_{\rm jk} } }} \right) \quad , \end{equation} replaces the second term on the r.h.s. of Eq.~(\ref{eq:en}). Here we see the effect of the nonlinear coupling in the ZM lagrangian. The strength of the scalar potential which is felt by particle j is weakened. We compare the Walecka forces with the ZM ones and study their effect on nuclear multifragmentation. The coupling strengths $g_{\rm s}$ and $g_{\rm v}$ are determined from the ground state properties of nuclear matter~\cite{Lindner,kn:serwal}. In order to include the symmetry energy we take also the $\varrho$ meson into account. To simplify the calculations the mass of the $\varrho$ meson is taken to be equal to the one of the $\omega$ meson, and the coupling constants are fitted to get the appropriate symmetry energy. In Table~1 we list all coupling constants used. \begin{table}[t] \begin{center} \begin{tabular}{||l|c|c|c|c|c||} \hline & $g_{\rm s}$ & $g_{\rm v}$ & $g_{\varrho}$ & $\mu_{\rm s} c^2$ & $\mu_{\rm v} c^2$ = $\mu_{\varrho} c^2$ \\ \hline \hline W & 11.04 & 13.74 & 7.0 & 550 MeV & 783 MeV \\ \hline ZM & 7.84 & 6.67 & 7.0 & 550 MeV & 783 MeV \\ \hline \end{tabular} \end{center} \caption{Parameters of the Walecka and the Zim\'anyi-Moszkowski forces} \end{table} Since the relativistic two-body forces are meant to describe the long range part of the interaction, the Yukawa functions are folded with a gaussian density distributions for the nucleons. For nuclear matter one gets a smooth density distribution if the gaussians $e^{-\alpha^2 (\vec{r} - \vec{r}_1 )^2}$ have a width parameter of $\alpha=0.5$ fm$^{-1}$. We use that value for folding the Yukawa forces, analogue to our earlier QMD calculations~\cite{kn:judqmd,kn:paula}, with $R^2_{\rm ij}$ replacing $\vec{r}^{\ 2}_{\rm ij}$ everywhere. \subsection*{\large\sl 3.2. Initial conditions} For the initial coordinate and momentum distribution of the particles we use the ones, which were prepared for the non-relativistic calculations~\cite{kn:judqmd} . The main aspect when generating initial positions and momenta is to get a smooth phase-space distribution. The ground-state energies calculated with distributions which fit nuclear charge densities are found to be good for both, the Walecka and the ZM forces. \subsection*{\large\sl 3.3. Cross sections} \begin{figure}[t] \vspace*{-10mm} \insertplot{ff-pdis.ps} \vspace*{-10mm} \caption{Distribution of nucleon-nucleon relative momenta up to time $t$ in Ca+Ca central collision} \end{figure} The multifragmentation depends on the cross sections used. According to Cugnon, his parameterization~\cite{kn:cug} gives a good fit only for relative momenta $\ge 0.8$ GeV/c in the two particle center of mass frame. For Ca + Ca central collisions at 200, 400, 600 and 800 MeV/u initial beam energies Fig.~2 shows however, that the vast majority of the collisions occur at relative momenta below $0.2$ GeV/c. To see the effect of the cross section on multifragmentation we fit the experimental free nucleon-nucleon cross sections~\cite{kn:land-b} as a function of the relative momentum with second order polynomials and use that fit in our calculations. However, due to the screening effect of the other nucleons we do not allow collisions at distances larger than $r_{\rm coll}$. The screening length is somewhat arbitrary, so we work with two parameterization, $r_{\rm coll}$ = 1.6 fm and $r_{\rm coll}$ = 2.4 fm. The low energy cross section, we use in the calculation, is displayed in Fig.~3. The dependence of the fragment distribution on the different cross sections is discussed in the next section. \begin{figure} \vspace*{-10mm} \insertplot{ff-cross.ps} \vspace*{-10mm} \caption{Low energy nucleon-nucleon cross section} \end{figure} \subsection*{\large\sl 3.4. Effective mass in two-body collisions} In relativistic many-body calculations single-particle energies $e_{\rm i}$ cannot be defined such that ($e_{\rm i},\vec{p}_{\rm i}$) form a 4-vector. This holds true only for the total momentum $\vec{P}$ and total energy $E$. However, we still try to deduce a single particle energy, which we use for prescribing the conservation of energy in each collision. If one considers only the scalar potential in the lowest order the solution is easy. The equation (\ref{eq:en}) in the leading order can be written as \begin{equation} \label{eq:efm} {E\over{mc^2}} = \sum_{\rm i}\ \sqrt{1+\vec{\tilde{p}}_{\rm i}^2} - {1\over 2} \sum_{\stackrel{{\scriptstyle \rm i,j}}{\rm j \neq i}}\ {\tilde{f}_{\rm ij} \over{ 1 + 2 {\displaystyle \sum_{\stackrel{{\scriptstyle \rm k}}{\rm k \neq j}}\ \tilde{f}_{\rm jk} } }} {1 \over{ \sqrt{1+\vec{\tilde{p}}_{\rm i}^2}}} + {1\over 2} \sum_{\stackrel{{\scriptstyle \rm ij}}{\rm j \neq i}}\ \tilde{g}_{\rm ji} \sqrt{1+\vec{\tilde{p}}_{\rm i}^2} \end{equation} There are two ways in our case to define the effective mass $m^*$ appearing in the collisions: Case A: If we neglect the momentum dependence in $\tilde{f}_{\rm ij}$ and $\tilde{g}_{\rm ij}$, the energy of the i$^{\rm th}$ particle can be written as \begin{eqnarray} \label{eq:sinen} {E(A) - E_{\rm i}(A-1) \over{ mc^2 }} &=& \sqrt{ 1 + \vec{\tilde{p}}_{\rm i}^2 } - {1 \over 2} \sum_{\stackrel{{\scriptstyle \rm j}}{\rm j \neq i}}\ {\tilde{f}_{\rm ij} \over { \sqrt{ 1 + \vec{\tilde{p}}_{\rm i}^2 } }} {1 \over{ 1 + 2 {\displaystyle \sum_{\stackrel{\scriptstyle \rm k}{\rm k \neq j}}\ \tilde{f}_{\rm jk} } }} \\ &+& {1 \over 2} \sum_{\stackrel{{\scriptstyle \rm j}}{\rm j \neq i}}\ \tilde{g}_{\rm ji} \sqrt{ 1 + \vec{\tilde{p}}_{\rm i}^2 } + \mbox{terms independent of }\ \vec{\tilde{p}}_{\rm i} \ ,\nonumber \end{eqnarray} where $E_{\rm i}(A-1)$ means the energy of the system when particle i is taken out. We can expand Eq.~(\ref{eq:sinen}) in $\vec{\tilde{p}}_{\rm i}^2$ and get $$ \epsilon_{\rm i} = {1 \over 2} \vec{\tilde{p}}_{\rm i}^2 \left[ 1 + {1 \over 2} \sum_{\stackrel{{\scriptstyle \rm j}}{\rm j \neq i}}\ {\tilde{f}_{\rm ij} \over{ 1 + 2 {\displaystyle \sum_{\stackrel{{\scriptstyle \rm k}}{\rm k \neq j}}\ \tilde{f}_{\rm jk} } }} + {1 \over 2} \sum_{\stackrel{{\scriptstyle \rm j}}{\rm j \neq i}}\ \tilde{g}_{\rm ji} \right] + \epsilon_{\rm i0} \quad , $$ where $\epsilon_{\rm i0}$ is independent of $\vec{\tilde{p}}_{\rm i}$. This can be written as $$ \epsilon_{\rm i} = \sqrt{ \left( m^*_{\rm i} \right)^2 + \vec{\tilde{p}}_{\rm i}^2 } + \epsilon^{\prime}_{\rm i0} \quad , $$ where the effective mass $m^*_{\rm i}$ turns out to be \begin{equation} \label{eq:msr} m^*_{\rm i} = 1 - {1 \over 2} \sum_{\stackrel{{\scriptstyle \rm j}}{\rm j \neq i}}\ {\tilde{f}_{\rm ij} \over{ 1 + 2 {\displaystyle \sum_{\stackrel{{\scriptstyle \rm k}}{\rm k \neq j}}\ \tilde{f}_{\rm jk} } }} - {1 \over 2} \sum_{\stackrel{{\scriptstyle \rm j}}{\rm j \neq i}}\ \tilde{g}_{\rm ji} \quad . \end{equation} \begin{figure}[t] \vspace*{-10mm} \insertplot{ff-fdti2.ps} \vspace*{-10mm} \caption{The evolution of the fragment distribution in time in 200 MeV/u O + Br central collisions} \end{figure} Case B: If we assume, that the momentum dependence of $\tilde{f}_{\rm ij}$ and $\tilde{g}_{\rm ji}$ is the same as in nuclear matter, that is \begin{equation} E = \sum_{\rm i}\ \sqrt{ 1 + \vec{\tilde{p}}_{\rm i}^2 } - {1\over 2} \sum_{\stackrel{{\scriptstyle \rm i,j}}{\rm j \neq i}}\ {f(r_{\rm ij}) \over{ \sqrt{ 1 + \vec{\tilde{p}}_{\rm i}^2 } \sqrt{ 1 + \vec{\tilde{p}}_{\rm j}^2 } }} { 1 \over { 1 + 2 {\displaystyle \sum_{\stackrel{{\scriptstyle \rm k}}{\rm k \neq j}}\ f(r_{\rm jk}) } }} + {1\over 2} \sum_{\stackrel{{\scriptstyle \rm i,j}}{\rm j \neq i}}\ g(r_{\rm ji}) \quad , \end{equation} (see Eq.~(\ref{eq:nms})),then the effective mass for finite systems turns out to be \begin{equation} \label{eq:msi} m^*_{\rm i} = 1 - \sum_{\stackrel{{\scriptstyle \rm j}}{\rm j \neq i}}\ {f(R_{\rm ij}) \over{ 1 + 2 {\displaystyle \sum_{\stackrel{{\scriptstyle \rm k}}{\rm k \neq j}}\ f(R_{\rm jk}) } }} \quad . \end{equation} We found in our calculations that the definition~(\ref{eq:msr}) gave better conservation of the total energy. \subsection*{\large\sl 3.5. The treatment of a relativistic collision} \begin{figure} \vspace*{-10mm} \insertplot{ff-fdti5.ps} \vspace*{-10mm} \caption{The evolution of the fragment distribution in time in 50 MeV/u O + Br central collision} \end{figure} The individual two-body collision is calculated in a relativistically covariant way, as given in \cite{kn:wolf}. The particles are allowed to collide if the impact parameter \begin{eqnarray} b &=& \sqrt{R_{12}^2 - h_{12}^2/v_{12}^2} \qquad \mbox{with} \\ R_{12}^2 &=& -(x_1 - x_2)^2 + \left( {p_1 (x_2 - x_1) \over{ m_1 c}} \right)^2 \quad , \nonumber \\ h_{12} &=& {p_1 (x_2 - x_1) \over{ m_1 c}} - {p_2 (x_2 - x_1) m_1 c \over{ p_1 p_2 }} \quad \mbox{and} \nonumber \\ v_{12}^2 &=& 1 - \left( {m_1 m_2 c^2 \over{ p_1 p_2 }} \right)^2 \nonumber \end{eqnarray} is smaller than the collision radius $\sqrt{\sigma/\pi}$. In the model the collision takes place at equal time in the CM of the total system, provided the closest approach of the world-lines occurs within the following time step. After the collision we only allow final momenta which do not let the particles approach each other further. \subsection*{\large\sl 3.6. Pauli blocking} Pauli blocking is treated as in Ref.~\cite{kn:judqmd}. We calculate the averaged phase-space distribution function at the location of each particle. The collision can occur if the new phase-space distribution function is smaller than 1 for the colliding particles and for all other particles after the collision. Since the distribution function at the phase space location of a particle depends on the momenta of all the other particles, without this condition it may occur that the phase space density becomes greater than one even for particles which did not take part in the collision. \subsection*{\large\sl 3.7. Definition of fragments} We consider the particles to belong to the same fragment if they stay together in coordinate and momentum space and are bound. In our calculation we follow the evolution of the system up to 360 fm/c for higher, and up to 720 fm/c for lower energies. We see in the Figs.~4 and 5 that in the last 120 fm/c the so determined fragment distribution does not change anymore. In order to calculate the energy of a cluster we calculate the $f_{\rm ij}$ and $g_{\rm ij}$ in the total CM frame because they are Lorentz-scalars. Only the momenta appearing in the momentum dependent terms have to be Lorentz transformed to the center of momentum frame of the cluster. The radii and the energies of the fragments are very near to the experimental values, although some fragments can shrink a little. However, we do not get energies lower than $-8.8$ MeV/u for any case. \begin{table}[t] \begin{center} \begin{tabular}{||c|c|c|c|c||} \hline & Scalar & Vector & $\vec{p}_{\rm i} \vec{p}_{\rm j}$ term & 3. order term \\ \hline \hline $\varrho_0$ & -153 & 122 & -0.01 & 2.2 \\ 1 MeV/u & -56 & 28 & 0.005 & 0.5 \\ \hline 3 $\varrho_0$ & -647 & 529 & 0.3 & 38 \\ 1 MeV/u & -90 & 59 & 0.04 & 2.19 \\ \hline $\varrho_0$ & -110 & 175 & -85 & 17 \\ 2 GeV/u & -40 & 40 & -19 & 14 \\ \hline 3 $\varrho_0$ & -235 & 381 & -169 & 80 \\ 2 GeV/u & -66 & 89.7 & -39 & 51.8 \\ \hline \end{tabular} \end{center} \caption{The contribution of the different energy terms of \protect{Eq.~(\ref{eq:en})} in MeV per particle for gold on gold systems for very extreme cases. The upper numbers refers to the Walecka, the lower ones to the ZM forces. The calculations are made for 1 MeV/u and 2 GeV/u energies at normal and 3 times normal nuclear matter density.} \end{table} \section*{\large\bf 4. Multifragmentation in the O\ +\ Br collision} Among the experiments carried out to study nuclear multifragmentation, emulsion measurements were the first to give an almost exclusive identification of the atomic numbers of the fragments emitted in heavy ion collisions. The investigation of the same system at several bombarding energies is particularly useful for testing models. Such data are available for the O + Br system at bombarding energies 50, 75, 100, 150 and 200 MeV/u \cite{kn:jakob}. Nowadays modern experimental techniques allow mass production of such measurements \cite{kn:FOPI}. Few years ago we carried out calculations for the emulsion experiment with a non-relativistic force to test the treatment of the Pauli blocking~\cite{kn:judqmd}. Now we have made the analogue calculation with relativistic forces at different energies to check their applicability at lower energies. These forces are meant to be used at higher energies, but they should also have good non-relativistic properties. \begin{figure} \vspace*{-10mm} \insertplot{ff-fdis.ps} \vspace*{-10mm} \caption{Fragment distribution in O + Br central collisions at energies 50, 75, 100, 150 and 200 MeV/u} \end{figure} As a first step we checked the validity of the expansion in $f_{\rm ij}$ and $g_{\rm ij}$. In Table~2 we give the values of different energy terms of Eq.~(\ref{eq:en}) for a gold on gold system which was artificially compressed to different densities and given different relative energies. One can see, that for the ZM case the contribution of the meson fields are much smaller than the nuclear rest mass $mc^2$. Even for the Walecka case, except maybe for the high compression with $j^0=3 \varrho_0$, the expansion is applicable. Thus our approximation in the expansion (\ref{eq:mom}) of the momentum is justified. In Fig.~6 the fragment distributions of the O + Br collisions are displayed for different energies. The contribution from the largest fragments are very small. Due to the computer time needed, the figures present results averaged over only 500 events. Thus the statistical fluctuation for events with small yields are still large. In Fig.~7 we compare the results obtained with the Walecka and ZM forces using the Cugnon parameterization \cite{kn:cug} and our fitted cross sections for 50 MeV/u and 200 MeV/u bombarding energies. The ZM force gives a somewhat better agreement with data than the forces derived from the Walecka lagrangian. The effect is larger for higher energies. However, using the Cugnon cross section gives more or less the same results as using the one fitted to the low energy free nucleon-nucleon scattering. The fragment distributions shown in Fig.~6 do not differ significantly from the earlier ones calculated with non-relativistic Skyrme forces, however one should mention, that for lower energies the experimental cut does not correspond to the central events. In Fig.~8 we display the distribution of impact parameters for the 10\% highest multiplicities. For the 200 MeV/u collision the multiplicity cut selects the central events (impact parameter is less than 3 fm), but for 50 MeV it results in broad distribution. The yields for 50 MeV/u and 200 MeV/u beam energies with the multiplicity cut are given in Fig.~9. \section*{\large\bf 5. Conclusion} \begin{figure} \vspace*{-10mm} \insertplot{ff-fdforc.ps} \vspace*{-10mm} \caption{Fragment distribution in O + Br central collisions at 200 MeV/u for different cross sections and forces} \end{figure} {}From an effective field theory lagrangian with scalar and vector mesons we deduced in the small acceleration approximation coordinate and momentum dependent relativistic forces, which are suitable for transport equation calculations. Even without an explicit density dependence, saturation of nuclear matter is obtained at proper densities and energies due to relativistic effects. The Zim\'anyi-Moszkowski forces, where a derivative coupling is used, can produce saturation even without relativistic effects. The proposed forces will be used in the intermediate energy domain (500 MeV/u -- 2 GeV/u), where the mean-field plays still an important role and relativistic forces are necessary. However, they should work at lower energies too. As a first check we studied their low energy behavior in a QMD calculation. We investigated both, the more static features of these forces in the ground-states of nuclei and their dynamical properties in the fragmentation process in heavy ion collisions. The fragment distributions in the O + Br (50 -- 200 A MeV) collision were calculated. The result we obtained is of similar quality as the one with the non-relativistic Skyrme forces~\cite{kn:judqmd,kn:paula}. That encourages us to use these forces for beam energies in the domain of 500 MeV/u -- 2 GeV/u, where non-relativistic Skyrme type forces cannot be used anymore. A further test of the model should be for example the comparison of the calculated flow (which is especially sensitive to the momentum dependence) for different energies with experimental results of the FOPI collaboration~\cite{kn:FOPI}. Such calculations are under progress. \begin{figure} \vspace*{-10mm} \insertplot{ff-bdis.ps} \vspace*{-10mm} \caption{The distribution of impact parameters to the highest multiplicities. The solid line corresponds to 50 MeV/u collisions and selects 11.3\% of the events, while the dashed line represents 9.5\% for 200 MeV/u} \end{figure} \begin{figure} \vspace*{-10mm} \insertplot{ff-50b.ps} \vspace*{-10mm} \insertplot{ff-200b.ps} \caption{The yields for 50 MeV/u (upper) and 200 MeV/u (lower) beam energies with trigger to the events with the 10\% highest multiplicities. The solid line displays the yield obtained for a central collision (b=0.5 fm), the dashed line displays the average over the highest multiplicity events. The experimental data are shown by crosses} \end{figure} \section*{\large\bf Acknowledgments} One of the authors (J.N.) should like to express her thanks to Prof. N\"orenberg for his kind hospitality at the GSI, where part of this work was done. G.P. thanks Gy. Wolf for fruitful discussions. This work was supported in part by the Hungarian Research Foundation OTKA. \section*{\large\bf Notes} E-mail: H. Feldmeier: <EMAIL> ; J. N\'e\-meth: <EMAIL> ;\\ G. Papp: <EMAIL>; WWW:http://www.gsi.de/\verb+~+papp .\\ $^\dagger$: Permanent address
\section{Introduction} The intention of this talk is to summarize the application of linearized gravity, in the specific form of the theory of black hole perturbations, to the study of the collision of black holes. Most of the results are already present in the literature, and the rest of the work is still in progress so I present here only a brief survey. The motivation for studying black hole collisions is quite clear. In the next few decades gravitational wave detectors will come online that will require ``templates'' of possible waveforms from different sources. The collision of black holes is one of the main candidates for observable sources of gravitational radiation. Although the initial and advanced LIGO detectors will not quite have the frequency range to detect the waves produced in the final moments of the most common collisions, it is expected that future detectors will, and knowing the waveform for the final moments can also lead to insights into the waveforms emitted earlier on. The presence of this strong motivation from the experimental side has led to the formation of an alliance of numerical relativity groups (the ``binary black hole grand challenge collaboration'') with the goal of numerically simulating the collision of two black holes using supercomputers. The degree of difficulty of this project is reflected in the fact that several established numerical relativity groups have decided to team efforts in order to tackle it. Here we will like to offer a much more modest approach, which is based on a simple idea: when a collision of two black holes starts with the holes so close to each other that they are surrounded by a common horizon, the problem looks from the point of view of an external observer as a single distorted black hole. It can therefore be treated with perturbation theory. Although one expects this approach to only yield results in a small range of initial separation, it provides ---at least for that range--- a benchmark against which one can calibrate numerical codes of the fully numerical approach. In reference \cite{PrPu} an explicit calculation was carried out using this idea. We took the initial data for the head-on collision of two black holes given by the Misner \cite{Mi} solution and re-wrote it in such a way that in the case that the two black holes are close to each other it explicitly looks like ``Schwarzschild plus something small". We took the ``something small" and evolved it using the equations of linearized gravity (the Zerilli equation) and computed the radiated energy. The results are shown in figure 1, where we plot the energy radiated in the collision as a function of the initial separation and compare with the results of the NCSA group \cite{NCSA} using a numerical integration of the full Einstein equations. We see that the close approximation works very well until the holes are no longer surrounded by a common apparent horizon ($\mu_0= 1.3$) and works within the correct order of magnitude up to when the holes are no longer surrounded by an event horizon ($\mu_0=2.0$). Also shown is a ``far approximation'' based on a particle-membrane paradigm \cite{collab}. Comparisons of waveforms have also been performed \cite{collab} and they also show very good agreement between the linearized theory and the full numerical simulations. \begin{figure} \vspace{-2.5cm} \epsfxsize=200pt \epsfbox{fig1.ps} \vspace{-0.5cm} \caption{Comparison of results for the radiated energy vs initial separation in a collision for the close approximation, the fully numerical results and the far approximation. Vertical scale is logarithmic. } \end{figure} All this shows that the use of linearized gravity in the close limit can be a valuable aid to full numerical evolutions of the two black hole problem. It is therefore quite tempting to apply the linearized treatment to more interesting situations, specifically the in-spiraling collision of two black holes with angular momentum. There are two main obstacles to doing this computation and we will detail them in the next two sections. \section{Second order perturbations: giving the formalism error bars} Assuming initial data for a black hole collision is given, we can rather easily evolve and compute energies in linearized theory. Why therefore not do it for the in-spiraling collision? The main reason is that for that case there are no numerical results with which to compare and the linearized formalism does not have a measure of error in it: it therefore has little predictive power. There is no consistent way to say when the close approximation breaks down. In fact, this example teaches us a valuable lesson about perturbation theory: when is linearized perturbation valid? The obvious answer ``when perturbations are small'' is clearly naive. To begin with, ``small'' should be characterized in a coordinate invariant way. Moreover, as this example shows, perturbations can be ``large'' and perturbation theory can still be valid: it just needs to happen that the perturbations be large in regions of spacetime that do not contribute in a significant way to the physics of interest. In the two black hole example, such a region is the interior of the horizon and regions close to it, in which perturbations mostly fall into the black hole. How is one to characterize when to trust the approximation? The answer is simple: work out the second order perturbations, compute the physical quantities of interest and use how much the first and second order calculation differ as a measure of the accuracy of the first order results. The advantage of this answer is that it is phrased in terms of what one is exactly interested in: the physical quantities. In the case of the collision of two black holes these are the radiated waveforms and energies. The formalism for second order perturbations of black holes has not been worked out in the past. It can be studied in detail as we do in reference \cite{zeri2}. Here I just sketch some of the outstanding points. It turns out that all the information can be coded into a single variable, exactly as in the first order perturbation case and that that variable satisfies a ``Zerilli equation'', \begin{equation} -{\partial \psi^{(2)} \over \partial t} + {\partial \psi^{(2)} \over \partial r_*} +V(r) \psi^{(2)} = S \end{equation} where $r_*=r+\log(r/2M-1)$ and the Zerilli function $\psi^{(2)}$ is a coordinate invariant combination of the perturbed metric coefficients. This equation is exactly the same as the one satisfied by the first order perturbations (including the ``potential'' V(r), which can be seen in reference \cite{PrPu}). However, there is an important difference: the right-hand side is not zero but a ``source'' term S, which is listed explicitly in reference \cite{zeri2} and which is a complicated function quadratic in the first order perturbations and their derivatives. The way in which we derived this equation is to compute a particular combination of the Einstein equations, writing the perturbed metric in a particular coordinate system, the so called ``Regge-Wheeler'' gauge. This, in turn is a way of deriving the original Zerilli equation. The expression we get for $\psi^{(2)}$ is therefore a representation in that gauge of a gauge invariant quantity. The explicitly gauge invariant form of $\psi^{(2)}$ can also be computed. We therefore are in a position to evolve to second order the problem of black hole collisions and therefore to endow the first order predictions with ``error bars''. This will be crucial for the inspiralling case, where numerical results are not expected for some time. \section{Initial data in the close approximation} In the head-on collision case we were lucky to have an exact solution to the initial value problem that we could evolve. For the more realistic cases there are no exact solutions available at present and it is unlikely that they will be easily found in the future. There is an immediate alternative at hand. There exist already well tested numerical codes \cite{Cook} for solving the initial value problem in general relativity in the context of black hole collisions. One could simply take these initial data evaluated for the case in which the black holes are close and ``read off'' from them the departures from Schwarzschild to be evolved using the linearized theory. This is certainly possible and has already been illustrated for Brill-Lindquist-type initial data by Abrahams and Price \cite{AbPr}. Apart from the possibility of using numerical initial data for realistic collisions it is interesting to notice that one can, up to a certain extent, solve the initial value problem analytically if one is only interested in initial data for the close approximation. The idea is simple: in the close approximation the initial data for a black hole collision departs a small amount from the initial data for a Schwarzschild spacetime for a single black hole with mass equal to the sum of the masses of the colliding holes. Therefore one can develop an approximation technique for the initial data starting from the initial data of Schwarzschild and adding small corrections proportional to the separation of the holes. We illustrate here only the zeroth order results, details will be given in a forthcoming paper in collaboration with John Baker. The initial value problem of general relativity can be conveniently cast in the conformal formalism \cite{Yo}. One is interested in solving the momentum and Hamiltonian constraints \begin{eqnarray} \nabla^a (K_{ab} - g_{ab} K) &=& 0\\ {}^3R-K_{ab} K^{ab} + K^2 &=&0 \end{eqnarray} where $g_{ab}$ is the spatial metric, $K_{ab}$ is the extrinsic curvature and ${}^3R$ is the scalar curvature of the three metric. One proposes a three metric that is conformally flat $g_{ab} = \psi^4 \delta_{ab}$, with $\psi^4$ the conformal factor and a decomposition of the extrinsic curvature $\widehat{K}_{ab} = \psi^{-2} K_{ab}$. The constraints become, \begin{eqnarray} \widehat{\nabla}^a \widehat{K}_{ab} &=& 0\\ \widehat{\nabla}^2 \psi &=& \psi^{-7} \widehat{K}_{ab} \widehat{K}^{ab}. \end{eqnarray} where $\widehat{\nabla}$ is a derivative with respect to the flat spacetime. Since the momentum constraint is linear, one can propose as a solution for it for the case of two black holes the sum of the solutions for the case of individual holes\footnote{The particular solution chosen depends on the boundary conditions imposed. This may add other terms to the simple ones we list here for brevity, but they all behave in a similar fashion with respect to the approximations we will consider.} with momentum $P_a$, \begin{equation} \hat{K}_{ab} = {3 \over 2 r^2} \left[ P_{(a} n_{b)} -(\delta_{ab} -n_a n_b)P^c n_c\right] \end{equation} where $n_{b}$ is a unit normal in the direction of $\vec{r}$ and all vector fields are defined in the flat background spacetime. One now can put this solution in the Hamiltonian constraint and one is left with an elliptic, highly non-linear equation for $\psi$. This is the equation that is usually solved numerically. There exist situations, however, where one can make some progress analytically. Consider the case in which the momenta of the holes is small \cite{Yo}. In that case one can neglect the right-hand side of the Hamiltonian constraint and one only needs to solve a vacuum Laplace equation for $\psi$. The solution can therefore be very simply found, the difficulty depending on the boundary conditions one chooses for the problem (typically a ``symmetrized'' boundary condition is imposed, which complicates calculation quite a bit in certain cases, see \cite{Cook} for details). Another situation in which one can obtain an approximate solution is in the ``close approximation''. In that case one has two black holes of momenta equal and opposite $P^{(1)}_a=-P^{(2)}_a$, and since the black holes are close, the unit normals appearing in the form for the extrinsic curvature for each hole are approximately equal. That implies that the extrinsic curvature for the problem is approximately zero (as it should, since in the close limit the problem looks like a Schwarzschild black hole at rest.) Therefore one can again neglect the right-hand side of the Hamiltonian constraint and one is again left with a Laplace equation. Let us compare this approximation with the full numerical results. In order to do this we will compare the ADM energy of initial data for a collision of two holes of momentum $P$. The ADM energy in the conformal formalism is given by \begin{equation} E=-{1\over 2 \pi}\oint_\infty \nabla_i\psi \, d^2 S^i \end{equation} and we notice that it does not depend explicitly on the extrinsic curvature (it does implicitly via the constraints). Therefore at the approximation we are working, in which the constraints do not couple the conformal factor and the extrinsic curvature, the energy is independent of the extrinsic curvature and therefore independent of the momenta of the holes. We compare this prediction with the full numerical results of Cook in figure \ref{cook}. \begin{figure} \vspace{-2.5cm} \epsfxsize=200pt \epsfbox{fig2.ps} \vspace{-0.5cm} \caption{The ADM energy of initial data for collisions of black holes of momenta $P$. The dots are the full numerical results of Cook, for different values of the initial separation $\beta$. We see that for small separations, the energy is approximately independent of the holes momenta, which coincides with the close approximation prediction, depicted by the solid line.} \label{cook} \end{figure} An interesting aspect is that one can advance this approximation one step further. One can input the extrinsic curvature and the conformal factor found as a fixed ``source'' in the equation determining the conformal factor and one can obtain a correction through the integration of a Poisson equation. Comparison of this approximation with the numerical data is currently in progress. Details are complicated by the particular boundary conditions that are usually chosen in the numerical computations. It is evident that the ``close approximation'' can work in many other cases, apart from the head-on, equal momenta holes we considered here. The only changes will be that the solution one obtains in the ``close limit'' rather than being a slice of Schwarzschild will be a slice of Kerr or boosted Schwarzschild if the net result of the collision has angular momentum or linear momentum. \section{Summary} We have seen that the use of the ``close approximation'' can be a valuable aid to full numerical computations of the collision of two black holes. With the introduction of a second order scheme we are now in a position of offering reliable estimates of energies and waveforms that we expect people working on the full numerical simulations will find of use to calibrate codes and design strategies for better integrating the Einstein equations in this problem of great current physical interest. \section*{Acknowledgments} The work described here is in collaboration with Richard Price, John Baker, Reinaldo Gleiser and Oscar Nicasio. I acknowledge support of NSF through grants PHY94-06269 PHY93-96246, funds of the Pennsylvania State University, its Office for Minority Faculty Development, and the Eberly Family research fund.
\section{\@startsection {section}{1}{\z@}{-2.5ex plus-1ex minus \newcommand{V_{ub}}{V_{ub}} \newcommand{V_{cb}}{V_{cb}} \newcommand{b\to u\ell\nu}{b\to u\ell\nu} \newcommand{b\to c\ell\nu}{b\to c\ell\nu} \newcommand{\Upsilon(4S)}{\Upsilon(4S)} \newcommand{GeV$/c$}{GeV$/c$} \newcommand{MeV$/c$}{MeV$/c$} \newcommand{\pi\ell\nu}{\pi\ell\nu} \newcommand{\rho\ell\nu}{\rho\ell\nu} \newcommand{\omega\ell\nu}{\omega\ell\nu} \newcommand{\rho^-\ell^+\nu}{\rho^-\ell^+\nu} \newcommand{\rho^0\ell^+\nu}{\rho^0\ell^+\nu} \newcommand{\omega\ell^+\nu}{\omega\ell^+\nu} \newcommand{\pi^-\ell^+\nu}{\pi^-\ell^+\nu} \newcommand{\pi^0\ell^+\nu}{\pi^0\ell^+\nu} \newcommand{\rho^\mp\ell^\pm\nu}{\rho^\mp\ell^\pm\nu} \newcommand{\rho^0\ell^\pm\nu}{\rho^0\ell^\pm\nu} \newcommand{\omega\ell^\pm\nu}{\omega\ell^\pm\nu} \newcommand{\pi^0\ell^\pm\nu}{\pi^0\ell^\pm\nu} \newcommand{\pi^\mp\ell^\pm\nu}{\pi^\mp\ell^\pm\nu} \newcommand{{\cal B}(B^0\to\pimlv)}{{\cal B}(B^0\to\pi^-\ell^+\nu)} \newcommand{{\cal B}(B^0\to\rhomlv)}{{\cal B}(B^0\to\rho^-\ell^+\nu)} \newcommand{\Gamma(B^0\to\pimlv)}{\Gamma(B^0\to\pi^-\ell^+\nu)} \newcommand{\Gamma(B^0\to\rhomlv)}{\Gamma(B^0\to\rho^-\ell^+\nu)} \newcommand{ B\to\pilv}{ B\to\pi\ell\nu} \newcommand{ D\to K\ell\nu}{ D\to K\ell\nu} \newcommand{ D\to\pilv}{ D\to\pi\ell\nu} \newcommand{ B\to\rholv}{ B\to\rho\ell\nu} \newcommand{ D\to\rholv}{ D\to\rho\ell\nu} \newcommand{ D\to K^{*}\ell\nu}{ D\to K^{*}\ell\nu} \newcommand{q^2_{max}}{q^2_{max}} \newcommand{E_{miss}}{E_{miss}} \newcommand{\vec{p}_{miss}}{\vec{p}_{miss}} \newcommand{\Delta E}{\Delta E} \newcommand{m_B}{m_B} \newcommand{\vec{p}_\nu}{\vec{p}_\nu} \newcommand{\vec{p}_\pi}{\vec{p}_\pi} \newcommand{\vec{p}_\ell}{\vec{p}_\ell} \newcommand{M_{miss}^2}{M_{miss}^2} \newcommand{N_{\pi^\pm\ell^\mp\nu}}{N_{\pi^\pm\ell^\mp\nu}} \newcommand{N_{\rho^\pm\ell^\mp\nu}}{N_{\rho^\pm\ell^\mp\nu}} \newcommand{\e}[1]{\times10^{#1}} \newcommand{\theta^*_{\pi\ell}}{\theta^*_{\pi\ell}} \newcommand{\cos\thepil}{\cos\theta^*_{\pi\ell}} \newcommand{fb$^{-1}$}{fb$^{-1}$} \newcommand{B\bar{B}}{B\bar{B}} \newcommand{{\it et al.}}{{\it et al.}} \newcommand{\plb}[1]{Phys. Lett. {\bf B#1}} \newcommand{\prl}[1]{Phys. Rev. Lett. {\bf #1}} \newcommand{\npb}[1]{Nucl. Phys. {\bf B#1}} \newcommand{\prd}[1]{Phys. Rev. {\bf D#1}} \newcommand{\zpc}[1]{Z. Phys. {\bf C#1}} \newcommand{\nim}[1]{Nucl. Instrum. Methods Phys. Res. Sect. A {\bf #1}} \newcommand{\begin{equation}}{\begin{equation}} \newcommand{\end{equation}}{\end{equation}} \newcommand{\hspace*{0.5em}}{\hspace*{0.5em}} \newcommand{\hspace*{0.5em}}{\hspace*{0.5em}} \renewcommand{\topfraction}{1.0} \renewcommand{\bottomfraction}{1.0} \renewcommand{\textfraction}{0.0} \renewcommand{\floatpagefraction}{0.75} \setcounter{totalnumber}{2} \begin{document} \begin{titlepage} \begin{center} \vspace*{1.0 truecm} {\elevenrm \hfill UR-1427}\\ {\elevenrm \hfill June, 1995}\\ \vspace*{2.5 truecm} {\twelvebf PRELIMINARY RESULTS FOR EXCLUSIVE \boldmath$b\to u\ell\nu$\\ DECAYS FROM CLEO}\\ \vspace{1 truecm} {\elevenrm Lawrence Gibbons}\\ {\elevenit University of Rochester, Rochester, NY 14627}\\ \newlength{\abskip} \setlength{\abskip}{4.0 truecm} \addtolength{\abskip}{-6\baselineskip} \vspace*{\abskip} \begin{abstract} { \noindent A preliminary analysis of exclusive $b\to u\ell\nu$ decays to the final states $\pi^\mp\ell^\pm\nu$, $\pi^0\ell^\pm\nu$, $\rho^\mp\ell^\pm\nu$, $\rho^0\ell^\pm\nu$ and $\omega\ell^\pm\nu$ based on $2.2\e{6}$ $B\bar{B}$ decays collected at CLEO is presented. We have measured the first exclusive $b\to u\ell\nu$ branching fraction ${\cal B}(B^0\to\pimlv)=[1.19\pm0.41\pm0.21\pm0.19]\e{-4}$ ($[1.70\pm0.51\pm0.31\pm0.27]\e{-4}$), with the ISGW (WSB) model used for efficiency determination. A 90\% C.L. upper limit on ${\cal B}(B^0\to\rhomlv)$ similar to the previous CLEO limit is obtained. The ratio $\Gamma(B^0\to\rhomlv)/\Gamma(B^0\to\pimlv)<3.4$ at the 90\% confidence level for both the ISGW and WSB models. This ratio provides some discrimination between form factor models. } \end{abstract} \vspace*{2.0 truecm} {\elevenit To appear in the Proceedings of the XXXth Rencontres de Moriond\\ ``Electroweak Interactions and Unified Theories'',\\ Les Arcs, France, March, 1995} \end{center} \end{titlepage} \section{Introduction} This talk will focus on a preliminary CLEO analysis of $b\to u\ell\nu$ decays to the exclusive final states $\pi\ell\nu$, $\rho\ell\nu$ and $\omega\ell\nu$. The ultimate goal of this analysis is to improve our knowledge of $|V_{ub}|$. ARGUS \cite{bb:argus_inclusive} and CLEO \cite{bb:cleo_inclusive} have already demonstrated that $|V_{ub}|>0$ by examining the inclusive lepton momentum spectrum from $B$ decays at the $\Upsilon(4S)$. They observe events beyond 2.4 GeV$/c$, which is kinematically forbidden for the copious $b\to c\ell\nu$ processes, but is still accessible to $b\to u\ell\nu$ decays. While these analyses clearly establish an excess in this endpoint region, and hence that $|V_{ub}|>0$, extracting a reliable value of $|V_{ub}|$ is difficult because of the theoretical uncertainty in extrapolating from the observed rate in the endpoint region to the total $b\to u\ell\nu$ rate. Values of $|V_{ub}/V_{cb}|$ obtained from these analyses are now in the 7\% to 11\% range, with the theoretical uncertainty dominating. \section{Exclusive $b\to u\ell\nu$} An alternate route to $|V_{ub}|$ is through the study of exclusive $b\to u\ell\nu$ channels. The best previous information concerning such channels is the upper limit set by CLEO \cite{bb:UCSBlimit} in the combined modes $\rho^-\ell^+\nu$, $\rho^0\ell^+\nu$ and $\omega\ell^+\nu$. The CLEO result corresponds to an upper limit of ${\cal B}(B^0\to\rhomlv)<3.2\e{-4}$ at the 90\% confidence level (ISGW model \cite{bb:isgw}). The preliminary analysis presented here studies the two pseudoscalar modes $\pi^-\ell^+\nu$ and $\pi^0\ell^+\nu$, the three vector modes $\rho^-\ell^+\nu$, $\rho^0\ell^+\nu$ and $\omega\ell^+\nu$, and the charge conjugate modes. At a fixed $|V_{ub}|$, the existing form factor models predict a wide range of partial widths for these modes, as Table~\ref{tab:exclpred} shows. Unfortunately, measured branching fractions depend on the form factor model used to evaluate the experimental efficiencies, as does the extraction of $|V_{ub}|$. We therefore need to discriminate between the different models. \begin{table}[b] \centering \caption{Predictions for the exclusive partial widths $\Gamma(B^0\to\pimlv)$ and $\Gamma(B^0\to\rhomlv)$ and the ratio $\Gamma(B^0\to\rhomlv)/\Gamma(B^0\to\pimlv)$. The partial width units are $10^{12}|V_{ub}|^2$ sec$^{-1}$.} \label{tab:exclpred} \begin{tabular}{lccc} \hline\hline Model & $\Gamma(B^0\to\pimlv)$ & $\Gamma(B^0\to\rhomlv)$ & $\Gamma(B^0\to\rhomlv)/\Gamma(B^0\to\pimlv)$\\ \hline WSB \cite{bb:WSB} & 6.3 -- 10.0 & 18.7 -- 42.5 & 3.0 -- 4.3 \\ KS \cite{bb:KS} & 7.25 & 33.0 & 4.6 \\ ISGW \cite{bb:isgw} & 2.1 & 8.3 & 4.0 \\ ISGW II \cite{bb:ISGWII} & 9.6 & 14.2 & 1.5 \\ \hline \end{tabular} \end{table} The ratio $\Gamma(B^0\to\rhomlv)/\Gamma(B^0\to\pimlv)$ provides one means of discrimination. Because the $\pi\ell\nu$ rate is helicity-suppressed when the daughter meson is at rest in the B meson rest frame (at $q^2_{max}$), where the form factors for the decay are largest, while the $\rho\ell\nu$ rate is not, we expect the ratio to be larger than one. The exact value for the ratio will depend on the $q^2$-behavior of the form factors. In Table~\ref{tab:exclpred}, we see that the predictions of the ratio span a fairly broad range, so the ratio should prove useful. \section{Neutrino ``Measurement'' and Exclusive $b\to u\ell\nu$} Experimentally, semileptonic decays are troublesome because of the undetected neutrino. This analysis takes advantage of the excellent hermeticity and resolution of the CLEO II detector located at the Cornell Electron Storage Ring (CESR) to obtain information about the neutrino in semileptonic $b\to u\ell\nu$ decays. Three concentric tracking devices provide a momentum resolution of $\sigma_p/p=0.005\oplus 0.0015p$ ($p$ in GeV$/c$), while covering 95\% of the $4\pi$ solid angle. The CsI calorimeter located inside of the CLEO solenoid provides an energy resolution well approximated by $\sigma_E/E=0.019+0.0035/E^{0.75}-0.001E$ ($E$ in GeV), while covering 98\% of $4\pi$. The detector is described in detail elsewhere \cite{bb:nim}. This analysis is based on a data sample with a luminosity of $2.09$ fb$^{-1}$\ (about $2.2\e{6}$ $B\bar{B}$ decays). The underlying idea is very simple: the $B\bar{B}$ system is at rest at CLEO and the beam energy is known very precisely, so we can ``measure'' the neutrino four momentum by measuring the missing energy and momentum of an event. We define \begin{eqnarray} E_{miss} & \equiv & 2E_{beam} - \sum_i E_i \\ \vec{p}_{miss} & \equiv & -\sum_i \vec{p}_i, \label{eq:missdef} \end{eqnarray} where the index $i$ runs over all charged tracks and all showers in the calorimeter that pass cuts designed to reject false tracks and spurious showers from hadronic interactions. In events with no extra missing particles, $\vec{p}_{miss}$ can be reliably associated with the momentum $\vec{p}_\nu$ of the signal mode neutrino. The $b\to u\ell\nu$ decay can then be fully reconstructed: the energy difference $\Delta E\equiv E_{beam} - (E_h+E_\ell+|\vec{p}_\nu|)$, where $h$ is the candidate hadron, should be zero, and the beam-energy constrained mass $m_B\equiv\sqrt{E_{beam}^2-|\vec{p}_h+\vec{p}_\ell+\vec{p}_\nu|^2}$ should reconstruct at the $B$ mass. Signal events that are reconstructible show resolutions of approximately 260 MeV on $E_{miss}$ and 110 MeV on $|\vec{p}_{miss}|$. Signal events with particles missing in addition to the neutrino usually fail the reconstruction criteria. On the other hand, those background events that pass the criteria do so because they have extra particles missing. Consequently, we reject events with multiple leptons or a non-zero total charge because they indicate a second neutrino or a missed charged particle, respectively. Most remaining events with extra missing particles are eliminated by requiring that $M_{miss}^2\equivE_{miss}^2 - |\vec{p}_{miss}|^2$ be consistent with zero. The criterion $M_{miss}^2/2E_{miss}<350$ MeV is used since the $M_{miss}^2$ resolution varies approximately as $2E_{miss}\sigma_{E_{miss}}$. Continuum background is suppressed using standard event shape variables. The $b\to u\ell\nu$ processes are enhanced over $b\to c$ by requiring the leptons to have momenta larger than 1.5 GeV$/c$\ (2.0 GeV$/c$) in the $\pi\ell\nu$ (vector) modes. The lower cut is used in the $\pi$ modes because these modes are expected to have a softer lepton momentum spectrum. Both electrons and muons are used in this analysis. We combine information from specific ionization, energy/momentum measurements from the calorimeter and tracking systems, and position matching from these two systems to identify electrons down to 600 MeV$/c$. Muon candidates must register hits in muon counters at least 5 interaction lengths deep, limiting the muon momentum range to approximately 1.4 GeV$/c$. The probability that a hadron is misidentified as a lepton (a ``fake lepton'') is of the order 0.1\% (1\%) for electron (muon) identification. Candidate $2\pi$ ($3\pi$) combinations must have an invariant mass within 90 (30) MeV of the nominal $\rho$ ($\omega$) mass. A $\pi^0$ candidate must have a 2-photon invariant mass within 2 standard deviations (about 12 MeV) of the $\pi^0$ mass. Within any one of the five modes, we pick the meson candidate in each event that yields the smallest value of $|\Delta E|$. We require the lepton, neutrino and meson candidates to satisfy $-250 \mbox{ MeV}<\Delta E<150\mbox{ MeV}$. The cut is asymmetric because the $b\to c$ backgrounds increase rapidly as $\Delta E$ increases. The range $5.265 \mbox{ GeV}<m_B<5.2875 \mbox{ GeV}$ defines the signal region. The $m_B$ distribution for data after all cuts, including the $\Delta E$ cut, is shown in Figure~\ref{fig:masses} for the combination of the $\pi^\mp\ell^\pm\nu$ and $\pi^0\ell^\pm\nu$ modes, and for the combination of the three vector modes. There is a clear excess above the background in the signal region for the $\pi\ell\nu$ modes. The fit yielding the background levels shown is described in the next section. The dominant background in both the $\pi$ and the vector modes comes from $b\to c\ell\nu$ decays in events containing either an undetected $K_L$ or a second neutrino. The small backgrounds in each mode from fake leptons and from continuum processes are measured with the data. In the $\pi$ modes, Monte Carlo studies indicate that feed-across from the $\rho\ell\nu$ modes should contribute the next largest background. In the vector modes, $b\to u\ell\nu$ decays to higher mass and non-resonant final states form the other major background component. Our fits do not make any requirement on the distribution of events inside our mass windows, so resonant and non-resonant final states are not distinguished. Consequently only an upper limit on ${\cal B}(B^0\to\rhomlv)$ will be obtained. We derive the limit conservatively by assuming zero background from the non-resonant and higher mass decays. \begin{figure}[t] \centering \leavevmode \epsfbox{morart_mass.eps} \caption{Beam constrained mass distributions for the combined $\pi^-\ell^+\nu$ and $\pi^0\ell^+\nu$ modes (left) and the combined vector modes (right). The points are continuum- and fake-subtracted data. The histograms show the contribution from $b\to c$ (shaded), $u\ell\nu$ crossfeed (hatched) and signal (hollow).} \label{fig:masses} \end{figure} \section{Extracting the Yields} \begin{table}[t] \centering \caption{Backgrounds, efficiencies and fit results for the $\pi\ell\nu$ analysis. The $\chi^2$ for the fits using the ISGW and WSB signal models were respectively 10.8 and 10.3 for $20-7$ degrees of freedom. Note that the errors on the signal yields and crossfeed backgrounds in the $\pi^-\ell^+\nu$ and $\pi^0\ell^+\nu$ modes are completely correlated because of the isospin constraints.} \label{tab:pifit} \begin{tabular}{lcccc} \hline\hline & \multicolumn{2}{c}{$\pi^-\ell^+\nu$} & \multicolumn{2}{c}{$\pi^0\ell^+\nu$} \\ \cline{2-5} & ISGW & WSB & ISGW & WSB \\ \hline Raw Data & \multicolumn{2}{c}{30} & \multicolumn{2}{c}{15} \\ Continuum Bkg. & \multicolumn{2}{c}{$2.3\pm0.8$} & \multicolumn{2}{c}{$1.0\pm0.5$} \\ Fake Lepton Bkg. & \multicolumn{2}{c}{$1.2\pm0.3$} & \multicolumn{2}{c}{$0.7\pm0.2$} \\ other $u\ell\nu$ Bkg. & \multicolumn{2}{c}{0.6} & \multicolumn{2}{c}{0.2} \\ Efficiency & 2.9\% & 2.1\% & 1.9\% & 1.4\% \\ Signal Yield & $15.6\pm5.3$ & $16.3\pm5.3$ & $5.0\pm1.7$ & $5.3\pm1.7$ \\ $b\to c$ Bkg. & $9.8\pm1.1$ & $9.8\pm1.1$ & $1.8\pm0.5$ & $1.7\pm0.5$ \\ $\rho/\omega$ Bkg.&$3.8\pm1.7$ & $3.4\pm1.4$ & $1.8\pm0.8$ & $1.6\pm0.7$ \\ \hline \end{tabular} \end{table} After subtracting the continuum and fake lepton backgrounds, we fit the beam-constrained mass distributions in our five reconstructed $b\to u\ell\nu$ modes simultaneously, which allows the data in the vector modes to constrain the $\rho\ell\nu$ background in the $\pi\ell\nu$ modes. In addition to the signal shapes and the feed-across shapes between the five modes, the fit includes $b\to c$ and other $b\to u\ell\nu$ background components. The isospin relations $\frac12\Gamma(B^0\to\pi^-\ell^+\nu) = \Gamma(B^+\to\pi^0\ell^+\nu)$ and $\frac12\Gamma(B^0\to\rho^-\ell^+\nu) = \Gamma(B^+\to\rho^0\ell^+\nu) \approx \Gamma(B^+\to\omega\ell^+\nu)$ constrain the neutral meson rates relative to the charged meson rates. We therefore obtain two yields, $N_{\pi^\pm\ell^\mp\nu}$ and $N_{\rho^\pm\ell^\mp\nu}$, from the fit. The $b\to c\ell\nu$ and feed-across background shapes in $m_B$ are obtained from Monte Carlo simulation. The $b\to c\ell\nu$ background level floats independently in each of the five modes, while the feed-across rates between the five modes are tied to the signal yields $N_{\pi^\pm\ell^\mp\nu}$ and $N_{\rho^\pm\ell^\mp\nu}$. Monte Carlo simulation also provides the $m_B$ distributions for the non-resonant and higher mass $b\to u\ell\nu$ backgrounds. The inclusive lepton yield at high momentum fixes this background level. We vary the physical model and the rate by hand to estimate the systematic uncertainty in this procedure. The results of the fit from which the $\pi\ell\nu$ yield (and the background levels in Figure~\ref{fig:masses}) is obtained are summarized in Table~\ref{tab:pifit}. The efficiencies and crossfeed probabilities have been determined using the ISGW and WSB models. We obtain similar $\pi\ell\nu$ yields for the two models, but obtain efficiencies that differ by approximately 30\%. The $b\to c\ell\nu$ background levels in the five modes are all consistent with absolute Monte Carlo predictions based on the luminosity. Correcting for acceptance and averaging the electron and muon samples, we obtain the preliminary branching fraction ${\cal B}(B^0\to\pimlv)=[1.19\pm0.41]\e{-4}$ ($[1.70\pm0.55]\e{-4}$) for the ISGW (WSB) model, where the errors are statistical only. We obtain consistent results if we fit using the $\Delta E$ distributions, having resolved multiple candidates using $m_B$. To obtain upper limits for the vector modes, we perform a similar fit assuming no non-resonant or higher mass $b\to u\ell\nu$ backgrounds. This fit gives the same $\pi\ell\nu$ yield. We obtain the efficiency-corrected numbers of $834\pm337$ ($1248\pm484$) $\rho^\mp\ell^\pm\nu$ decays for the ISGW (WSB) model, and a $\Gamma(B^0\to\rhomlv)/\Gamma(B^0\to\pimlv)$ ratio of $1.56^{+1.29}_{-0.76}$ ($1.63^{+1.21}_{-0.75}$). \begin{figure}[b] \centering \leavevmode \epsfbox{morart_lepthe.eps} \caption{Charged lepton spectrum (left) and $\cos\thepil$ distribution (right) for the combined $\pi^-\ell^+\nu$ and $\pi^0\ell^+\nu$ modes. The points are continuum- and fake-subtracted data. The top histogram is the total prediction using rates from the yield fit, with components $b\to c$ (shaded), $u\ell\nu$ crossfeed (hatched) and signal (dashed).} \label{fig:lepspec} \end{figure} Many distributions have been examined for consistency with the $\pi\ell\nu$ hypothesis. The charged lepton momentum spectrum for $\pi^\mp\ell^\pm\nu$ and $\pi^0\ell^\pm\nu$ candidates in the $m_B$ signal region is shown in Figure~\ref{fig:lepspec}. The spectrum obtained from the data is quite stiff, with a sizeable fraction of events beyond the $b\to c\ell\nu$ endpoint. The sum of the signal and background distributions, scaled according to the fit results, shows good agreement with the data. The $\pi$ and $\nu$ momentum spectra are also consistent with the results of the fit. For $B\to\pi\ell\nu$, the $V-A$ interaction predicts that the angle between the $\pi$ and the lepton in the $W$ rest frame, $\theta^*_{\pi\ell}$, should have a $\sin^2\theta^*_{\pi\ell}$ distribution. The observed $\cos\thepil$ distribution, also shown in Figure~\ref{fig:lepspec}, is in good agreement with this expectation. We estimate the probability, including systematic uncertainties, that the background processes could fluctuate to give the observed $m_B$ and $\cos\thepil$ distributions in the combined $\pi\ell\nu$ modes, and obtain $6.4\e{-5}$. This corresponds to a 3.8 standard deviation significance for a Gaussian distribution. \section{Systematics} \begin{table}[t] \centering \caption{Summary of systematic uncertainties on the yields and efficiencies in the $\pi\ell\nu$ and $\rho\ell\nu$ modes. The numbers in parentheses in the background levels indicate the uncertainty in the background as a fraction of that background.} \label{tab:syst} \begin{tabular}{lrr||lrrr} \hline\hline On yields: & $\pi\ell\nu$ & $\rho\ell\nu$ & On Efficiencies: & $\pi\ell\nu$ & $\rho\ell\nu$ & $\rho/\pi$ ratio \\ \hline $b\to c$ bkg. & (20\%) 13\% & (20\%) 23\% & $\nu$-measurement & 15\% & 15\% & 15\% \\ $\rho/\omega\ell\nu$ bkg.& (36\%)\hspace*{0.5em}\ 8\%& (63\%)\hspace*{0.5em}\ 7\% & $\pi/\rho/\omega$ finding& 3\% & 6\% & 7\% \\ other $u\ell\nu$ bkg. & 8\% & --- & $\rho/\omega$ polarization & --- & 10\% & 10\% \\ cont.+fake bkg. & (20\%)\hspace*{0.5em}\ 6\%& (24\%)\hspace*{0.5em}\ 7\% & lepton fake rates & 4\% & 4\% & 4\% \\ lepton finding & 2\% & 2\% & lepton finding & 4\% & 4\% & 4\% \\ & & & Luminosity & 2\% & 2\% & --- \\ \hline {\bf Total} & {\bf 18\%} & {\bf 25\%} & {\bf Total} & {\bf 16\%} & {\bf 20\%} & {\bf 20\%} \\ \hline \end{tabular} \end{table} \begin{figure}[b] \centering \leavevmode \epsfbox{morart_dstar.eps} \caption{$|\vec{p}_{miss}|$ spectrum (left) $m_B$ distribution (center) and $\Delta E$ distribution (right) for $D^{*\pm}\ell^\mp\nu$ reconstruction. The points are continuum and combinatoric background-subtracted data. The histograms are signal Monte Carlo distributions normalized to equal area.} \label{fig:dstar} \end{figure} The systematic uncertainties on the yields and efficiencies are summarized in Table~\ref{tab:syst}. The dominant uncertainty in the yields comes from the uncertainty in the shapes of the background $m_B$ distributions. The shapes have been checked in a variety of ways: examining the shapes in $\Delta E$ sidebands and in signal-free modes (eg., $K_S\ell\nu$), and varying the misreconstruction behavior of the Monte Carlo simulation. The uncertainty in the efficiencies is dominated by the neutrino-measurement simulation. One method of estimating this uncertainty is to use this technique to measure the branching fraction for $B\to D^{*\pm}\ell^\mp\nu$ via the modes $D^{*\pm}\to\pi^\pm D^0$, $D^0\to K^\mp\pi^\pm$. We find that the simulation of $E_{miss}$, $\vec{p}_{miss}$, $\Delta E$ and $m_B$ agrees well with the data (Figure~\ref{fig:dstar}), and that ${\cal B}(B^0\to D^{*-}\ell^+\nu)=4.66\pm0.65\%$. This agrees with the published CLEO result \cite{bb:dstar} of $4.49\pm0.32\pm0.32\%$, which used a higher statistics technique. The 15\% statistical uncertainty is taken as the systematic uncertainty; other studies indicate that this is a conservative estimate. We expect this systematic to cancel in the $\rho/\pi$ ratio, but retain a preliminary 15\% uncertainty. \section{Conclusion} Combining the $\pi\ell\nu$ yields and the systematic uncertainties, we obtain the preliminary branching fraction ${\cal B}(B^0\to\pi^-\ell^+\nu)=[1.19\pm0.41\pm0.21\pm0.19]\e{-4}$ ($[1.70\pm0.51\pm0.31\pm0.27]\e{-4}$) using the ISGW (WSB) model to evaluate efficiencies. The errors are statistical, systematic on the yield, and systematic on the efficiency, respectively. This is the first measurement of any exclusive $b\to u\ell\nu$ branching fraction. The probability of a background fluctuation resulting in the observed signal is $6.4\e{-5}$. Assuming no non-resonant or high mass $u\ell\nu$ background, we obtain a conservative 90\% C.L. upper limit of ${\cal B}(B^0\to\rho^-\ell^+\nu) < 3.1\e{-4}$ for the ISGW model and ${\cal B}(B^0\to\rho^-\ell^+\nu) < 4.6\e{-4}$ for the WSB model. The statistical and systematic uncertainties have been combined in quadrature in evaluating these limits. The results are comparable to the previous CLEO upper limits for the vector modes. Finally, we find $\Gamma(B^0\to\rhomlv)/\Gamma(B^0\to\pimlv)<3.4$ at the 90\% confidence level for both the ISGW and WSB models. Again, statistical and systematic uncertainties have been combined in quadrature. Comparing to the predictions in Table~\ref{tab:exclpred}, the WSB model is compatible with this limit, but it is difficult to reconcile the ISGW model with this limit. These preliminary measurements herald a new era for the study of $V_{ub}$. CLEO is still refining these measurements, with 50\% more data soon to be available and work in progress on the separation of the vector modes from non-resonant modes. \section{References}
\section{Introduction} \label{secintro} The description of small-$x$\ partons within hadrons has attracted a great deal of interest, especially after the measurements of the proton structure function $F_2$ at $x$\ values down to $10^{-4}$ at HERA, where a substantial increase was found \cite{H1struct93,ZEUSstruct93}. Although the small-$x$\ rise was first predicted by the so-called BFKL evolution equation \cite{BFKL77:1,BFKL77:2} it soon turned out \cite{BallForte941,BallForte942} that it could also be explained in terms of the conventional Altarelli--Parisi (DGLAP) evolution equations \cite{DGLAP:1,DGLAP:2,DGLAP:3,DGLAP:4}. Instead, much of the focus has been directed towards the study of hadronic final states in deep inelastic lepton--hadron scattering (DIS) events at small $x$, and it has been suggested that the large flow of transverse energy in the proton direction found in such events is a signal of BFKL dynamics \cite{MARTINetflow94}. Much can also be learned from comparing data with different models implemented in Monte Carlo event generators. So far it has been shown that generators built around a conventional DGLAP-inspired initial-state parton showers, such as \pythia\ \cite{JETPYT94,JETPYT93} and \lepto\ \cite{LEPTO91}, with strong ordering in virtuality, completely fail to describe things like the forward transverse energy flow at small $x$, while a generator such as \ariadne\ \cite{ARIADNENOW} -- although not implementing BFKL evolution, but sharing with it the feature that emissions are unordered in transverse momenta -- describes such event features quite well \cite{H1flow94,H1flow95}. Besides deep inelastic lepton--hadron scattering, Drell--Yan\ production in hadron--hadron collisions is one of the cleanest probes of hadronic structure. Recent results from the D0 collaboration \cite{GEOFFcdm94} at the Tevatron shows a surprising feature of events with Drell--Yan-produced $W$ bosons, namely the decorrelation in rapidity between the $W$ and the associated jets. Although the typical $x$-values probed in $W$ events at the Tevatron is on the order of {\small $\sqrt{m_W^2/S}$} $\approx 80/1800 \approx 0.04$, for large rapidities of the $W$, one of the incoming partons has a much smaller momentum fraction of the proton (e.g.\ $y_W \approx 2$ gives $x_1 \approx 6 \times 10^{-3}$ and $x_2 \approx 0.3$). Therefore it could be worth while to take the experience gained from studying small-$x$\ final states at HERA and try to apply it to large rapidity $W$-production at the Tevatron. In this paper the Dipole Cascade Model (DCM) \cite{CDMinit86,CDMplain88}, on which the \ariadne\ program is built, is extended to also model the jet structure of Drell--Yan\ production events. The main feature of the DCM for DIS \cite{CDMdis89} is that gluon emission is treated as final state radiation from the colour dipole formed between the struck quark and the proton remnant as in fig.\refig{figdisdip}a. In this way there is no explicit initial state radiation, and the proton structure enters only in the way the dipole radiation is suppressed due to the spatial extension of the proton remnant. This approach has some problems when it comes to describing features particular to the initial state, such as the initial splitting of a gluon into a $\qq$-pair. \begin{figure} \setlength{\unitlength}{0.07mm} \begin{picture}(2000,600)(-100,0) \thicklines \put(200,125){\line(1,0){600}} \thinlines \put(200,150){\line(1,0){200}} \put(400,150){\vector(1,2){50}} \put(400,150){\line(1,2){100}} \put(500,350){\line(2,1){300}} \put(500,350){\vector(2,1){150}} \qbezier(500,350)(450,350)(450,390) \qbezier(450,390)(450,430)(400,430) \qbezier(400,430)(350,430)(350,470) \qbezier(350,470)(350,510)(300,510) \qbezier(300,510)(250,510)(250,550) \qbezier(250,550)(250,590)(200,590) \put(450,550){\makebox(0,0){$\gamma/Z^0$}} \qbezier(850,450)(900,300)(850,150) \put(850,450){\vector(-1,4){0}} \put(850,150){\vector(-1,-4){0}} \thicklines \put(1200,125){\line(1,0){600}} \put(1200,575){\line(1,0){600}} \thinlines \put(1200,150){\line(1,0){200}} \put(1400,150){\vector(1,2){50}} \put(1400,150){\line(1,2){100}} \put(1500,350){\line(-1,2){100}} \put(1500,350){\vector(-1,2){50}} \put(1400,550){\line(-1,0){200}} \multiput(1500,350)(50,0){5}{\line(1,0){25}} \put(1650,300){\makebox(0,0){$W$}} \qbezier(1825,150)(1875,350)(1825,550) \put(1825,150){\vector(-1,-4){0}} \put(1825,550){\vector(-1,4){0}} \put(400,50){\makebox(0,0){(a)}} \put(1400,50){\makebox(0,0){(b)}} \end{picture} \fcap{The colour dipoles that initiate the dipole cascade in (a) DIS and (b) Drell--Yan\ production of $W$.} \label{figdisdip} \end{figure} The simplest extension of the DCM to also treat Drell--Yan\ production would be to treat gluon emission as final-state radiation from the colour dipole formed between the two remnants, as in fig.\refig{figdisdip}b. However, as is seen from fig.\refig{figordalp}, the leading order $W$+jet diagrams all correspond to initial-state radiation (except for the last one, which is the least important). And in particular it is clear that if the gluon emission is treated as final-state radiation between the two remnants, it would be difficult to explain the contribution to the transverse momentum of the $W$ from the diagrams in fig.\refig{figordalp}. \begin{figure} \setlength{\unitlength}{0.07mm} \begin{picture}(2000,600)(-100,0) \put(0,50){\line(1,1){150}} \put(150,200){\line(0,1){200}} \put(150,400){\line(-1,1){150}} \multiput(150,400)(40,40){4}{\begin{picture}(70,70) \qbezier(0,0)(30,30)(60,0) \qbezier(60,0)(70,-10)(60,-20) \qbezier(60,-20)(50,-30)(40,-20) \qbezier(40,-20)(10,10)(40,40)\end{picture}} \multiput(150,200)(40,-40){4}{\begin{picture}(20,20) \qbezier(0,0)(10,-10)(20,-20)\end{picture}} \put(0,550){\vector(1,-1){75}} \put(150,400){\vector(0,-1){100}} \put(150,200){\vector(-1,-1){75}} \put(355,300){\makebox(0,0){{\large +}}} \put(355,20){\makebox(0,0){(a)}} \put(400,50){\line(1,1){150}} \put(550,200){\line(0,1){200}} \put(550,400){\line(-1,1){150}} \multiput(550,200)(40,-40){4}{\begin{picture}(70,70) \qbezier(0,0)(30,-30)(60,0) \qbezier(60,0)(70,10)(60,20) \qbezier(60,20)(50,30)(40,20) \qbezier(40,20)(10,-10)(40,-40)\end{picture}} \multiput(550,400)(40,40){4}{\begin{picture}(20,20) \qbezier(0,0)(10,10)(20,20)\end{picture}} \put(400,550){\vector(1,-1){75}} \put(550,400){\vector(0,-1){100}} \put(550,200){\vector(-1,-1){75}} \multiput(1000,40)(40,40){4}{\begin{picture}(70,70) \qbezier(0,0)(30,30)(60,0) \qbezier(60,0)(70,-10)(60,-20) \qbezier(60,-20)(50,-30)(40,-20) \qbezier(40,-20)(10,10)(40,40)\end{picture}} \put(1160,200){\line(0,1){200}} \multiput(1160,400)(40,40){4}{\begin{picture}(20,20) \qbezier(0,0)(10,10)(20,20)\end{picture}} \put(1160,400){\line(-1,1){160}} \put(1320,40){\line(-1,1){160}} \put(1000,560){\vector(1,-1){80}} \put(1160,400){\vector(0,-1){100}} \put(1160,200){\vector(1,-1){80}} \put(1370,300){\makebox(0,0){{\large +}}} \put(1370,20){\makebox(0,0){(b)}} \multiput(1420,140)(40,40){4}{\begin{picture}(70,70) \qbezier(0,0)(30,30)(60,0) \qbezier(60,0)(70,-10)(60,-20) \qbezier(60,-20)(50,-30)(40,-20) \qbezier(40,-20)(10,10)(40,40)\end{picture}} \put(1580,300){\line(-1,1){160}} \put(1780,300){\line(-1,0){200}} \multiput(1780,310)(40,40){4}{\begin{picture}(20,20) \qbezier(0,0)(10,10)(20,20)\end{picture}} \put(1940,140){\line(-1,1){160}} \put(1420,460){\vector(1,-1){80}} \put(1580,300){\vector(1,0){100}} \put(1780,300){\vector(1,-1){80}} \end{picture} \fcap{The leading-order Feynman diagrams contributing to $W$ + jet production corresponding to the annihilation (a) and Compton (b) diagrams.} \label{figordalp} \end{figure} In previous work \cite{CDMdispom94} the DCM was extended to also include the boson--gluon fusion diagram in DIS, which can be viewed as a special case of initial-state gluon splitting into $\qq$. In this paper, this approach is further developed into a general inclusion of initial-state gluon splitting into $\qq$. Also, a way is presented of taking into account the contribution to the transverse momentum of the $W$ from the gluon emission, formulated it terms of radiation from the colour dipole between the two proton remnants. The layout of this paper is as follows. In sections \ref{secglu} and \ref{seccmp} the treatment of gluon emission and initial state gluon splitting is presented. In section \ref{secres}, results for the $W$--jet rapidity correlation from the improved DCM is compared to a leading-order calculation and with the conventional DGLAP-inspired initial-state parton shower approach of \pythia. Also some predictions are given for the transverse energy flow in high rapidity $W$ events at Tevatron energies. Finally, in section \ref{secsum}, the conclusions are presented. \section{Gluon Emission} \label{secglu} The DCM for $\ee$\ annihilation and deep inelastic lepton--hadron scattering is described in detail in refs.\ \cite{CDMinit86,CDMplain88,CDMdis89} and only a brief summary of the features important for this paper will be given here. The emission of a gluon $g_1$ from a $\qq$\ pair created in an $\ee$\ annihilation event can be described as radiation from the colour dipole between the $q$ and $\bar{q}$. A subsequent emission of a softer gluon $g_2$ can be described as radiation from two independent colour dipoles, one between the $q$ and $g_1$ and one between $g_1$ and $\bar{q}$. Further gluon emissions are given by three independent dipoles etc. In DIS, the gluon emission comes from the dipole stretched between the quark, struck by the electro--weak probe, and the hadron remnant. The situation is the same as in $\ee$\ above, except that, while $q$ and $\bar{q}$\ are both point-like in the case of $\ee$, the hadron remnant in DIS is an extended object. In an antenna of size $l$, radiation with wavelengths $\lambda \ll l$ are strongly suppressed. In the DCM, this is taken into account by only letting a fraction \begin{equation} a=\mu/k_\perp \label{eqremfrac1} \end{equation} of the hadron remnant take part in the emission of a gluon with transverse momentum $\kt$, where $\mu$ is a parameter corresponding to the inverse (transverse) size of the hadron. The phase space available in dipole emission is conveniently pictured by the inside of a triangle in the $\kappa$--$y$ plane, where $\kappa\equiv\ln{k_\perp^2}$ and $y$ is the rapidity of the emitted gluon as in fig.\refig{figphase1}. In these variables the dipole emission cross section also takes a particularly simple approximate form: \begin{equation} d\sigma\propto\alpha_Sd\kappa dy. \label{eqMEdip1} \end{equation} In DIS, assuming that the hadron is coming in with momentum (using light-cone coordinates) ($P_+,0,\vec{0})$, and is probed by a virtual photon $(-Q_+,Q_-,\vec{0})$, the triangular area comes from the trivial requirement \begin{eqnarray} k_{+g} & \equiv & k_\perp e^y < P_+\nonumber\\ k_{-g} & \equiv & k_\perp e^{-y} < Q_-. \end{eqnarray} The condition that only a fraction of the remnant participates in an emission means that \begin{equation} k_{+g}<(\mu/k_\perp)P_+ \label{eqphasecut} \end{equation} and translates into an extra cutoff in the phase space corresponding to the thick line in fig.\refig{figphase1}. This should be compared to the initial-state parton shower scenario, where gluon emission is given by \begin{equation} d\sigma_{q}=\frac{2\alpha_S}{3\pi} \frac{1+z^2}{1-z}\frac{f_q(x/z)}{f_q(x)}\frac{dz}{z}\frac{dQ^2}{Q^2}. \label{eqiniqsplit1} \end{equation} Identifying the ratio of structure functions in eq.\req{eqiniqsplit1} (corresponding to the dotted line of equal suppression in fig.\refig{figphase1}) with the extra cutoff (\ref{eqphasecut}) in the DCM, the two models are equivalent in the low-$\kt$\ limit. \begin{figure} \setlength{\unitlength}{0.07mm} \begin{picture}(2000,700)(-100,0) \put(1050,650){\makebox(0,0){$\kappa$}} \put(1650,100){\makebox(0,0){$y$}} \put(600,50){\vector(-1,0){50}} \put(625,50){\makebox(0,0)[l]{{\footnotesize direction of struck quark}}} \put(400,100){\vector(1,0){1200}} \put(1000,100){\vector(0,1){600}} \put(500,100){\line(1,1){500}} \put(1500,100){\line(-1,1){500}} \thicklines \put(1520,100){\line(-2,1){800}} \qbezier[40](1300,100)(1100,300)(900,500) \end{picture} \fcap{The phase space available for gluon emission in DIS (thin lines) and the extra restriction due to the extendedness of the proton remnant (thick line). The dotted line corresponds to a line of equal suppression due to the ratio of parton density functions entering into a conventional parton shower scenario.} \label{figphase1} \end{figure} As mentioned in the introduction, the simplest way of extending the DCM to describe gluon emissions in Drell--Yan\ events is to describe it as radiation from the colour dipole between the two hadron remnants. One problem with this approach is what to do with the transverse recoil from the gluon emission. In $\ee$, this recoil is shared by the $q$ and $\bar{q}$. In DIS, since only a fraction of the remnant is taking part in the emission, only that fraction is given a transverse recoil, resulting in an extra, so-called recoil gluon \cite{CDMdis89}. The corresponding procedure for Drell--Yan\ would be to introduce two recoil gluons, one for each remnant. However in that way it is impossible to reproduce the transverse momentum of the $W$ as given by the \tordas\ matrix element. The \tordas\ matrix element for $q+\bar{q}\rightarrow W+g$ production takes the form \cite{Halzen78} \begin{equation} M^{q\bar{q}\rightarrow Wg} \propto \frac{\hat{t}^2+\hat{u}^2+2m_W^2\hat{s}}{\hat{t}\hat{u}}, \label{eqMEWg1} \end{equation} where $\hat{s}$, $\hat{t}$ and $\hat{u}$ are the ordinary Mandelstam variables satisfying $\hat{s}+\hat{t}+\hat{u}=m_W^2$. In order to reproduce this in a parton shower scenario, where the gluon is emitted ``after'' the $W$ is produced, this has to be convoluted with the parton density functions, and the lowest-order $W$-production matrix element, again convoluted with the relevant parton densities, has to be factored out. This introduces some ambiguities, which are solved by assuming that the rapidity of the $W$ is the same before and after the gluon emission, resulting in the following cross section, expressed in the transverse momentum $\k2t$\ and rapidity $y_g$ of the emitted gluon \begin{eqnarray} \frac{d\sigma_g}{dy_gdk_\perp^2} & = & \frac{2\alpha_S}{3\pi} \frac{f_q(x_q')}{f_q(x_q)} \frac{f_{\bar{q}}(x_{\bar{q}}')}{f_{\bar{q}}(x_{\bar{q}})} \times \nonumber \\ & & \frac{(k_\perp^2+m_\perp^2+k_\perpm_\perp e^{\Delta y})^2+(k_\perp^2+m_\perp^2+k_\perpm_\perp e^{-\Delta y})^2}{(k_\perp^2+k_\perpm_\perp e^{\Delta y})(k_\perp^2+k_\perpm_\perp e^{-\Delta y})(k_\perp^2+m_\perp^2+k_\perpm_\perp(e^{\Delta y}+e^{-\Delta y}))}, \label{eqMEWg2} \end{eqnarray} where $\Delta y = y_g-y_W$, $m_\perp^2=k_\perp^2+m_W^2$, $y_W$ the rapidity of the $W$ and $x_i$ and $x_i'$ the energy--momentum fractions carried by the incoming partons before and after the gluon emission so that \begin{eqnarray} x_q = \frac{m_W e^{y_W}}{\sqrt{S}}, & & x_q' = \frac{m_\perp e^{y_W} + k_\perp e^{y_g}}{\sqrt{S}}, \\ x_{\bar{q}} = \frac{m_W e^{-y_W}}{\sqrt{S}}, & & x_{\bar{q}}' = \frac{m_\perp e^{-y_W} + k_\perp e^{-y_g}}{\sqrt{S}}, \end{eqnarray} assuming the $q$ coming in along the positive $z$-axis and a total invariant mass of $\sqrt{S}$. In the limit $k_\perp^2\ll m_W^2$, eq.\req{eqMEWg2} reduces to the simple dipole emission cross section in eq.\req{eqMEdip1}, so it is clear that the strategy outlined above is a good leading log approximation. It also turns out that it is fairly simple to correct the first gluon emission so that, disregarding the ratios of parton densities, eq.\req{eqMEWg2} is reproduced. \begin{figure} \setlength{\unitlength}{0.07mm} \begin{picture}(2000,700)(-100,0) \put(1050,650){\makebox(0,0){$\kappa$}} \put(1650,100){\makebox(0,0){$y$}} \put(400,100){\vector(1,0){1200}} \put(1000,100){\vector(0,1){600}} \put(500,100){\line(1,1){500}} \put(1500,100){\line(-1,1){500}} \thicklines \put(1520,100){\line(-2,1){800}} \put(480,100){\line(2,1){800}} \put(1100,50){\vector(0,1){50}} \put(1100,25){\makebox(0,0){$y_W$}} \put(1400,400){\vector(-1,0){50}} \put(1420,400){\makebox(0,0)[l]{$\ln{m_W^2}$}} \put(600,350){\vector(1,0){50}} \put(575,350){\makebox(0,0)[r]{$\ln{k_{\perp\max}^2}$}} \qbezier[31](800,100)(950,250)(1100,400) \qbezier[30](820,100)(965,245)(1110,390) \qbezier[29](840,100)(980,240)(1120,380) \qbezier[28](860,100)(995,235)(1130,370) \qbezier[27](880,100)(1010,230)(1140,360) \qbezier[26](900,100)(1025,225)(1150,350) \qbezier[25](920,100)(1040,220)(1160,340) \qbezier[24](940,100)(1055,215)(1170,330) \qbezier[23](960,100)(1070,210)(1180,320) \qbezier[22](980,100)(1085,205)(1190,310) \qbezier[21](1000,100)(1100,200)(1200,300) \qbezier[20](1020,100)(1115,195)(1210,290) \qbezier[19](1040,100)(1130,190)(1220,280) \qbezier[18](1060,100)(1145,185)(1230,270) \qbezier[17](1080,100)(1160,180)(1240,260) \qbezier[16](1100,100)(1175,175)(1250,250) \qbezier[15](1120,100)(1190,170)(1260,240) \qbezier[14](1140,100)(1205,165)(1270,230) \qbezier[13](1160,100)(1220,160)(1280,220) \qbezier[12](1180,100)(1235,155)(1290,210) \qbezier[11](1200,100)(1250,150)(1300,200) \qbezier[10](1220,100)(1265,145)(1310,190) \qbezier[9](1240,100)(1280,140)(1320,180) \qbezier[8](1260,100)(1295,135)(1330,170) \qbezier[7](1280,100)(1310,130)(1340,160) \qbezier[6](1300,100)(1325,125)(1350,150) \qbezier[5](1320,100)(1340,120)(1360,140) \qbezier[4](1340,100)(1355,115)(1370,130) \qbezier[3](1360,100)(1370,110)(1380,120) \qbezier[2](1380,100)(1385,105)(1390,110) \end{picture} \fcap{The phase space available for gluon emission in $W$ production (thin lines) and the extra restriction due to the extendedness of the proton remnants (thick lines). The shaded triangle corresponds to the phase space area covered by the $W$.} \label{figphase2} \end{figure} The ratio of parton densities in eq.\req{eqMEWg2} is instead approximated by the suppression of the phase space introduced for DIS in eq.\req{eqphasecut}, which in this case corresponds to suppressions on both sides of the triangle, as in fig.\refig{figphase2}. One problem with this procedure is that the $\k2t$\ of the gluon and hence of the $W$ is limited by this suppression to \begin{equation} k_\perp^2<\mu\sqrt{S/4}, \end{equation} which, with $\mu\approx 1$ GeV, gives $k_\perp\,\lower3pt\hbox{$\buildrel < \over\sim$}\, 30$ GeV. To be able to describe high-$\kt$\ $W$ production, it is clear that the sharp cutoff in fig.\refig{figphase2}, which in any case is an oversimplification, must be replaced by a smooth suppression. In \cite{CDMdis89} it was shown that introducing a power suppression in the disallowed regions in fig.\refig{figphase1} does not influence the general event shapes in DIS; it is clear, however, that such a power suppression would greatly influence the high-$\kt$\ spectrum of the $W$. First, however, the way of obtaining a transverse momentum of the $W$ in the gluon emissions must be formalized. It is clear that, in the first emission, the gluon corresponds unambiguously to initial state radiation, and hence the $\kt$\ of the gluon must be balanced by the $\kt$\ of the $W$. In further emissions this is not the case, as the dipole radiation is a coherent sum of the emission from the incoming partons and the outgoing, previously radiated, gluon. It is therefore argued that only gluon radiation that takes place close to the $W$ in phase space should be able to influence the $\kt$\ of the $W$; only gluons emitted in the shaded region of fig.\refig{figphase2}, corresponding to $P_{g+}<P_{W+}$ and $P_{g-}<P_{W-}$, will have their transverse recoil absorbed by the $W$. Outside this region the transverse recoil will be treated as in the DIS case above. \begin{figure} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2880,1728)(-500,-100) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 0 1426 V LTb 600 251 M 63 0 V 2034 0 R -63 0 V 600 394 M 31 0 V 2066 0 R -31 0 V 600 478 M 31 0 V 2066 0 R -31 0 V 600 537 M 31 0 V 2066 0 R -31 0 V 600 583 M 31 0 V 2066 0 R -31 0 V 600 621 M 31 0 V 2066 0 R -31 0 V 600 653 M 31 0 V 2066 0 R -31 0 V 600 680 M 31 0 V 2066 0 R -31 0 V 600 705 M 31 0 V 2066 0 R -31 0 V 600 726 M 63 0 V 2034 0 R -63 0 V 600 869 M 31 0 V 2066 0 R -31 0 V 600 953 M 31 0 V 2066 0 R -31 0 V 600 1013 M 31 0 V 2066 0 R -31 0 V 600 1059 M 31 0 V 2066 0 R -31 0 V 600 1096 M 31 0 V 2066 0 R -31 0 V 600 1128 M 31 0 V 2066 0 R -31 0 V 600 1156 M 31 0 V 2066 0 R -31 0 V 600 1180 M 31 0 V 2066 0 R -31 0 V 600 1202 M 63 0 V 2034 0 R -63 0 V 600 1345 M 31 0 V 2066 0 R -31 0 V 600 1428 M 31 0 V 2066 0 R -31 0 V 600 1488 M 31 0 V 2066 0 R -31 0 V 600 1534 M 31 0 V 2066 0 R -31 0 V 600 1572 M 31 0 V 2066 0 R -31 0 V 600 1603 M 31 0 V 2066 0 R -31 0 V 600 1631 M 31 0 V 2066 0 R -31 0 V 600 1655 M 31 0 V 2066 0 R -31 0 V 600 1677 M 63 0 V 2034 0 R -63 0 V 600 251 M 0 63 V 0 1363 R 0 -63 V 862 251 M 0 63 V 0 1363 R 0 -63 V 1124 251 M 0 63 V 0 1363 R 0 -63 V 1386 251 M 0 63 V 0 1363 R 0 -63 V 1649 251 M 0 63 V 0 1363 R 0 -63 V 1911 251 M 0 63 V 0 1363 R 0 -63 V 2173 251 M 0 63 V 0 1363 R 0 -63 V 2435 251 M 0 63 V 0 1363 R 0 -63 V 2697 251 M 0 63 V 0 1363 R 0 -63 V 600 251 M 2097 0 V 0 1426 V -2097 0 V 600 251 L LT0 2394 1514 M 180 0 V 810 251 M 0 447 V 52 0 V 0 495 V 53 0 V 0 -52 V 52 0 V 0 -56 V 52 0 V 0 -49 V 53 0 V 0 -53 V 52 0 V 0 -45 V 53 0 V 0 -44 V 52 0 V 0 -51 V 53 0 V 0 -32 V 52 0 V 0 -42 V 52 0 V 0 -25 V 53 0 V 0 -49 V 52 0 V 0 -25 V 53 0 V 0 -33 V 52 0 V 0 -49 V 53 0 V 0 6 V 52 0 V 0 -39 V 52 0 V 0 -43 V 53 0 V 0 -54 V 52 0 V 0 -28 V 53 0 V 0 -9 V 52 0 V 0 -8 V 52 0 V 0 -47 V 53 0 V 0 -37 V 52 0 V 0 7 V 53 0 V 0 -50 V 52 0 V 0 33 V 53 0 V 0 -68 V LT1 2394 1414 M 180 0 V 600 1551 M 52 0 V 0 80 V 53 0 V 0 -76 V 52 0 V 0 -104 V 53 0 V 0 -113 V 52 0 V 0 -111 V 53 0 V 0 -123 V 52 0 V 0 -122 V 52 0 V 0 -124 V 53 0 V 0 -130 V 52 0 V 0 -123 V 53 0 V 0 -143 V 52 0 V 0 -80 V 53 0 V 0 -131 V LT3 2394 1314 M 180 0 V 600 1563 M 52 0 V 0 40 V 53 0 V 0 -91 V 52 0 V 0 -99 V 53 0 V 0 -94 V 52 0 V 0 -81 V 53 0 V 0 -74 V 52 0 V 0 -68 V 52 0 V 0 -62 V 53 0 V 0 -59 V 52 0 V 0 -54 V 53 0 V 0 -46 V 52 0 V 0 -42 V 53 0 V 0 -55 V 52 0 V 0 -33 V 52 0 V 0 -26 V 53 0 V 0 -59 V 52 0 V 0 -41 V 53 0 V 0 -36 V 52 0 V 0 -24 V 53 0 V 0 -33 V 52 0 V 0 -32 V 52 0 V 0 -29 V 53 0 V 0 -37 V 52 0 V 0 -16 V 53 0 V 0 -51 V 52 0 V 0 -16 V 52 0 V 0 -18 V 53 0 V 0 -7 V 52 0 V 0 -41 V 53 0 V 0 -4 V 52 0 V 0 -24 V stroke grestore end showpage } \put(2334,1314){\makebox(0,0)[r]{DCM gluons $\beta=2$}} \put(2334,1414){\makebox(0,0)[r]{DCM gluons $\beta=\infty$}} \put(2334,1514){\makebox(0,0)[r]{LO $W+g$}} \put(1648,-49){\makebox(0,0){$k_{\perp W}$ (GeV)}} \put(220,964){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$d\sigma/dk_{\perp W}$ (pb/GeV)}}% \special{ps: currentpoint grestore moveto}% } \put(2697,151){\makebox(0,0){80}} \put(2435,151){\makebox(0,0){70}} \put(2173,151){\makebox(0,0){60}} \put(1911,151){\makebox(0,0){50}} \put(1649,151){\makebox(0,0){40}} \put(1386,151){\makebox(0,0){30}} \put(1124,151){\makebox(0,0){20}} \put(862,151){\makebox(0,0){10}} \put(600,151){\makebox(0,0){0}} \put(540,1677){\makebox(0,0)[r]{1000}} \put(540,1202){\makebox(0,0)[r]{100}} \put(540,726){\makebox(0,0)[r]{10}} \put(540,251){\makebox(0,0)[r]{1}} \end{picture} \fcap{The transverse momentum spectrum of the $W$ at the Tevatron. The full line is the prediction of the \tordas\ $W+g$ matrix element as implemented in \pythia\ using CTEQ2L parton density functions. The matrix element calculation was cut off at $k_{\perp W}=10$ GeV to avoid divergences. The dashed and dotted lines are the predictions of the gluon emission in the DCM with $\beta=\infty$ and $\beta=2$, respectively.} \label{figptWg} \end{figure} Figure \ref{figptWg} shows the \tordas\ $W+g$ matrix element prediction (as implemented in \pythia) of the $\kt$\ spectrum of the $W$ at the Tevatron, compared to the modified DCM described above. With a sharp cutoff in the phase space, it is clear that the DCM cannot describe the high-$\kt$\ tail of the spectrum. Instead, a smooth suppression can be introduced by changing eq.\req{eqremfrac1}, allowing a larger fraction $a'$ of the remnant to take part in the emission with the probability \begin{equation} P(a') \propto \frac{\beta(\frac{a'}{a})^\beta}{a'(1+(\frac{a'}{a})^\beta)^2}, \end{equation} corresponding to a smoothening of the theta function suppression in fig.\refig{figphase2}, giving a power-suppressed tail. As seen in fig.\refig{figptWg}, using $\beta=2$ describes well the high-$\kt$\ spectrum obtained from the leading-order calculation using the CTEQ2L\footnote{The CTEQ2L structure function parametrization is used in all analyses in this paper where applicable. None of the conclusions in this paper were found to be sensitive to this choice.} \cite{CTEQ2} structure function parametrization. In the following, this value of $\beta$ will be used, unless stated otherwise. This concludes the description of the $W$ + gluon jet in the DCM. However, at small $x$, the gluon density in the proton becomes very large, and the Compton diagrams in fig.\refig{figordalp}b are dominating. \section{The Compton Diagrams} \label{seccmp} The matrix element for the Compton diagrams looks like \cite{Halzen78} \begin{equation} M^{gq\rightarrow Wq} \propto -\frac{\hat{s}^2+\hat{t}^2+2m_W^2\hat{u}}{\hat{s}\hat{t}}. \label{eqMEWq1} \end{equation} The $s$-channel diagram is, of course heavily suppressed for small $\kt$, and looking only at the $t$-channel diagram, convoluting with parton densities and factoring out the zeroth order $W$ production cross section, as in the gluon emission case above, the normal leading log initial-state parton shower cross section for the splitting of an incoming gluon into a $\qq$\ pair \cite{TSinips88} is obtained: \begin{equation} d\sigma_{q}=\frac{\alpha_S}{4\pi} (z^2+(1-z)^2)\frac{f_g(x/z)}{f_q(x)}\frac{dz}{z}\frac{dQ^2}{Q^2}, \label{eqinigsplit1} \end{equation} where $Q^2=-\hat{t}$ and $z=m_W^2/\hat{s}$. In the DCM, however, there is no initial state gluon splitting into $\qq$, and, just as in the case of final-state $\g2qq$\ splitting \cite{CDMsplit90}, this process has to be added by hand to the DCM. The simplest way is to introduce the initial-state $\g2qq$\ splitting in the same way as in \cite{CDMsplit90}, as a process competing with the DCM gluon emission described above. The competition is as usual governed by the Sudakov form factor using ordering in $\k2t$. Rewriting eq.\req{eqinigsplit1} in terms of the transverse momentum $\k2t$\ and rapidity $y_q$ of the outgoing quark, the probability of the {\em first} emission to be an initial-state $\g2qq$\ splitting at a certain $\k2t$\ and $y_q$ is given by \begin{eqnarray} \frac{dP_{q}(k_\perp^2,y_q)}{dk_\perp^2 dy_q} & = & \frac{d\sigma_{q}(k_\perp^2,y_q)}{dk_\perp^2 dy_q} \times \nonumber \\ & & \exp{-\int_{k_\perp^2}^{k_{\perp\max}^2}dk_\perp^{2\prime}\int dy_q' \left(\frac{d\sigma_{q}(k_\perp^{2\prime},y_q')}{dk_\perp^{2\prime} dy_q'}+\frac{d\sigma_{g}(k_\perp^{2\prime},y_q')}{dk_\perp^{2\prime} dy_q'}\right)}, \end{eqnarray} where the second factor is the Sudakov form factor, corresponding to the probability {\em not} to have any emission of gluons {\em or} gluon splittings above the scale $\k2t$. Technically, the extra process is implemented as follows. If the quark going into the hard interaction on one side is a sea-quark, the remnant on that side is allowed to ``radiate'' the corresponding antiquark according to eq.\req{eqinigsplit1}. After such an emission, the remnant is split in two parts according to the prescription described in ref.\ \cite{CDMdis89}, one of which forms a dipole with the ``emitted'' antiquark while the other retains the dipole colour connection of the original remnant. If the first emission is a $\g2qq$\ splitting, the full \tordas\ matrix element is used, and the rapidity of the $W$ is assumed to be the same before and after the emission, as in the gluon emission case. A splitting later on in the cascade, the kinematic is fixed by requiring the non-radiating remnant to be unchanged. In all cases, the transverse momentum of the struck system will of course balance the $\kt$\ of the emitted antiquark. Note that only one initial state $\g2qq$\ splitting is allowed per remnant. This is a good approximation, since a second such splitting is heavily suppressed by the parton density functions. This procedure can be used not only in the case of $W$ production, but for all processes with a hadron remnant present. In particular it can be (and is\footnote{This is the default in \ariadne\ version 4.06 and later.}) applied in the DIS case. \begin{figure} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2880,1728)(-500,-400) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 0 1326 V LTb 600 251 M 63 0 V 2034 0 R -63 0 V 600 351 M 31 0 V 2066 0 R -31 0 V 600 409 M 31 0 V 2066 0 R -31 0 V 600 451 M 31 0 V 2066 0 R -31 0 V 600 483 M 31 0 V 2066 0 R -31 0 V 600 509 M 31 0 V 2066 0 R -31 0 V 600 531 M 31 0 V 2066 0 R -31 0 V 600 550 M 31 0 V 2066 0 R -31 0 V 600 567 M 31 0 V 2066 0 R -31 0 V 600 583 M 63 0 V 2034 0 R -63 0 V 600 682 M 31 0 V 2066 0 R -31 0 V 600 741 M 31 0 V 2066 0 R -31 0 V 600 782 M 31 0 V 2066 0 R -31 0 V 600 814 M 31 0 V 2066 0 R -31 0 V 600 840 M 31 0 V 2066 0 R -31 0 V 600 863 M 31 0 V 2066 0 R -31 0 V 600 882 M 31 0 V 2066 0 R -31 0 V 600 899 M 31 0 V 2066 0 R -31 0 V 600 914 M 63 0 V 2034 0 R -63 0 V 600 1014 M 31 0 V 2066 0 R -31 0 V 600 1072 M 31 0 V 2066 0 R -31 0 V 600 1114 M 31 0 V 2066 0 R -31 0 V 600 1146 M 31 0 V 2066 0 R -31 0 V 600 1172 M 31 0 V 2066 0 R -31 0 V 600 1194 M 31 0 V 2066 0 R -31 0 V 600 1213 M 31 0 V 2066 0 R -31 0 V 600 1230 M 31 0 V 2066 0 R -31 0 V 600 1246 M 63 0 V 2034 0 R -63 0 V 600 1345 M 31 0 V 2066 0 R -31 0 V 600 1404 M 31 0 V 2066 0 R -31 0 V 600 1445 M 31 0 V 2066 0 R -31 0 V 600 1477 M 31 0 V 2066 0 R -31 0 V 600 1503 M 31 0 V 2066 0 R -31 0 V 600 1526 M 31 0 V 2066 0 R -31 0 V 600 1545 M 31 0 V 2066 0 R -31 0 V 600 1562 M 31 0 V 2066 0 R -31 0 V 600 1577 M 63 0 V 2034 0 R -63 0 V 600 251 M 0 63 V 0 1263 R 0 -63 V 1019 251 M 0 63 V 0 1263 R 0 -63 V 1439 251 M 0 63 V 0 1263 R 0 -63 V 1858 251 M 0 63 V 0 1263 R 0 -63 V 2278 251 M 0 63 V 0 1263 R 0 -63 V 2697 251 M 0 63 V 0 1263 R 0 -63 V 600 251 M 2097 0 V 0 1326 V -2097 0 V 600 251 L LT0 2394 1414 M 180 0 V 768 251 M 0 686 V 42 0 V 0 354 V 42 0 V 0 -29 V 42 0 V 0 -38 V 42 0 V 0 -30 V 41 0 V 0 -35 V 42 0 V 0 -23 V 42 0 V 0 -33 V 42 0 V 0 -25 V 42 0 V 0 -28 V 42 0 V 0 -16 V 42 0 V 0 -23 V 42 0 V 0 -21 V 42 0 V 0 -23 V 42 0 V 0 -29 V 42 0 V 0 -16 V 42 0 V 0 -15 V 42 0 V 0 -30 V 42 0 V 42 0 V 0 -24 V 42 0 V 0 -18 V 42 0 V 0 -22 V 41 0 V 0 -12 V 42 0 V 0 -3 V 42 0 V 0 -26 V 42 0 V 0 -29 V 42 0 V 0 -7 V 42 0 V 0 -12 V 42 0 V 0 -33 V 42 0 V 0 -8 V 42 0 V 0 -8 V 42 0 V 0 -16 V 42 0 V 0 -25 V 42 0 V 0 -1 V 42 0 V 0 -32 V 42 0 V 0 -18 V 42 0 V 0 -8 V 42 0 V 0 -24 V 41 0 V 0 7 V 42 0 V 0 5 V 42 0 V 0 -40 V 42 0 V 0 -19 V 42 0 V 0 -8 V 42 0 V 0 -25 V 42 0 V 42 0 V 0 10 V 42 0 V 0 -37 V LT3 2394 1314 M 180 0 V 600 1437 M 42 0 V 0 50 V 42 0 V 0 -41 V 42 0 V 0 -46 V 42 0 V 0 -44 V 42 0 V 0 -46 V 42 0 V 0 -40 V 42 0 V 0 -35 V 42 0 V 0 -36 V 41 0 V 0 -28 V 42 0 V 0 -24 V 42 0 V 0 -32 V 42 0 V 0 -34 V 42 0 V 0 -14 V 42 0 V 0 -18 V 42 0 V 0 -31 V 42 0 V 0 -11 V 42 0 V 0 -36 V 42 0 V 0 -12 V 42 0 V 0 -27 V 42 0 V 0 -24 V 42 0 V 0 -27 V 42 0 V 0 2 V 42 0 V 0 -40 V 42 0 V 0 -2 V 42 0 V 0 -45 V 41 0 V 0 8 V 42 0 V 0 -15 V 42 0 V 0 -20 V 42 0 V 0 -12 V 42 0 V 0 -42 V 42 0 V 0 30 V 42 0 V 0 -42 V 42 0 V 0 -8 V 42 0 V 0 -7 V 42 0 V 0 -53 V 42 0 V 0 -12 V 42 0 V 0 36 V 42 0 V 0 -53 V 42 0 V 0 -3 V 42 0 V 0 7 V 42 0 V 0 -75 V 41 0 V 0 -12 V 42 0 V 0 69 V 42 0 V 42 0 V 0 -82 V 42 0 V 0 -93 V 42 0 V 0 63 V 42 0 V 0 -9 V 42 0 V 0 -54 V 42 0 V 0 100 V stroke grestore end showpage } \put(2334,1314){\makebox(0,0)[r]{Full DCM}} \put(2334,1414){\makebox(0,0)[r]{Leading order}} \put(1648,1677){\makebox(0,0){(a)}} \put(1648,-49){\makebox(0,0){$k_{\perp W}$ (GeV)}} \put(280,914){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$d\sigma/dk_{\perp W}$ (pb/GeV)}}% \special{ps: currentpoint grestore moveto}% } \put(2697,151){\makebox(0,0){100}} \put(2278,151){\makebox(0,0){80}} \put(1858,151){\makebox(0,0){60}} \put(1439,151){\makebox(0,0){40}} \put(1019,151){\makebox(0,0){20}} \put(600,151){\makebox(0,0){0}} \put(540,1577){\makebox(0,0)[r]{1000}} \put(540,1246){\makebox(0,0)[r]{100}} \put(540,914){\makebox(0,0)[r]{10}} \put(540,583){\makebox(0,0)[r]{1}} \put(540,251){\makebox(0,0)[r]{0.1}} \end{picture} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2880,1728)(-500,-100) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 0 1326 V LTb 600 251 M 63 0 V 2034 0 R -63 0 V 600 351 M 31 0 V 2066 0 R -31 0 V 600 409 M 31 0 V 2066 0 R -31 0 V 600 451 M 31 0 V 2066 0 R -31 0 V 600 483 M 31 0 V 2066 0 R -31 0 V 600 509 M 31 0 V 2066 0 R -31 0 V 600 531 M 31 0 V 2066 0 R -31 0 V 600 550 M 31 0 V 2066 0 R -31 0 V 600 567 M 31 0 V 2066 0 R -31 0 V 600 583 M 63 0 V 2034 0 R -63 0 V 600 682 M 31 0 V 2066 0 R -31 0 V 600 741 M 31 0 V 2066 0 R -31 0 V 600 782 M 31 0 V 2066 0 R -31 0 V 600 814 M 31 0 V 2066 0 R -31 0 V 600 840 M 31 0 V 2066 0 R -31 0 V 600 863 M 31 0 V 2066 0 R -31 0 V 600 882 M 31 0 V 2066 0 R -31 0 V 600 899 M 31 0 V 2066 0 R -31 0 V 600 914 M 63 0 V 2034 0 R -63 0 V 600 1014 M 31 0 V 2066 0 R -31 0 V 600 1072 M 31 0 V 2066 0 R -31 0 V 600 1114 M 31 0 V 2066 0 R -31 0 V 600 1146 M 31 0 V 2066 0 R -31 0 V 600 1172 M 31 0 V 2066 0 R -31 0 V 600 1194 M 31 0 V 2066 0 R -31 0 V 600 1213 M 31 0 V 2066 0 R -31 0 V 600 1230 M 31 0 V 2066 0 R -31 0 V 600 1246 M 63 0 V 2034 0 R -63 0 V 600 1345 M 31 0 V 2066 0 R -31 0 V 600 1404 M 31 0 V 2066 0 R -31 0 V 600 1445 M 31 0 V 2066 0 R -31 0 V 600 1477 M 31 0 V 2066 0 R -31 0 V 600 1503 M 31 0 V 2066 0 R -31 0 V 600 1526 M 31 0 V 2066 0 R -31 0 V 600 1545 M 31 0 V 2066 0 R -31 0 V 600 1562 M 31 0 V 2066 0 R -31 0 V 600 1577 M 63 0 V 2034 0 R -63 0 V 600 251 M 0 63 V 0 1263 R 0 -63 V 1019 251 M 0 63 V 0 1263 R 0 -63 V 1439 251 M 0 63 V 0 1263 R 0 -63 V 1858 251 M 0 63 V 0 1263 R 0 -63 V 2278 251 M 0 63 V 0 1263 R 0 -63 V 2697 251 M 0 63 V 0 1263 R 0 -63 V 600 251 M 2097 0 V 0 1326 V -2097 0 V 600 251 L LT0 2394 1414 M 180 0 V 600 678 M 42 0 V 0 202 V 42 0 V 0 106 V 42 0 V 0 90 V 42 0 V 0 81 V 42 0 V 0 46 V 42 0 V 0 7 V 42 0 V 0 -16 V 42 0 V 0 -19 V 41 0 V 0 -28 V 42 0 V 0 -18 V 42 0 V 0 -24 V 42 0 V 0 -23 V 42 0 V 0 -19 V 42 0 V 0 -22 V 42 0 V 0 -24 V 42 0 V 0 -14 V 42 0 V 0 -22 V 42 0 V 0 -12 V 42 0 V 0 -28 V 42 0 V 0 -14 V 42 0 V 0 -35 V 42 0 V 0 7 V 42 0 V 0 -49 V 42 0 V 0 19 V 42 0 V 0 -18 V 41 0 V 0 -29 V 42 0 V 0 5 V 42 0 V 0 -8 V 42 0 V 0 -49 V 42 0 V 0 2 V 42 0 V 0 -23 V 42 0 V 0 -8 V 42 0 V 0 -17 V 42 0 V 0 -34 V 42 0 V 0 -7 V 42 0 V 0 -19 V 42 0 V 42 0 V 0 -71 V 42 0 V 0 53 V 42 0 V 0 -53 V 42 0 V 0 -20 V 41 0 V 0 33 V 42 0 V 0 -100 V 42 0 V 0 31 V 42 0 V 0 25 V 42 0 V 0 -106 V 42 0 V 0 58 V 42 0 V 0 -58 V 42 0 V 0 12 V 42 0 V 0 -38 V LT1 2394 1314 M 180 0 V 600 1504 M 42 0 V 0 -19 V 42 0 V 0 -57 V 42 0 V 0 -53 V 42 0 V 0 -42 V 42 0 V 0 -35 V 42 0 V 0 -35 V 42 0 V 0 -37 V 42 0 V 0 -30 V 41 0 V 0 -27 V 42 0 V 0 -33 V 42 0 V 0 -18 V 42 0 V 0 -33 V 42 0 V 0 -29 V 42 0 V 0 -21 V 42 0 V 0 -25 V 42 0 V 0 -15 V 42 0 V 0 -33 V 42 0 V 0 -27 V 42 0 V 0 -21 V 42 0 V 0 -22 V 42 0 V 0 -21 V 42 0 V 0 -18 V 42 0 V 0 -38 V 42 0 V 0 -5 V 42 0 V 0 -43 V 41 0 V 0 -33 V 42 0 V 0 -43 V 42 0 V 0 -38 V 42 0 V 0 -8 V 42 0 V 42 0 V 0 -67 V 42 0 V 0 -13 V 42 0 V 0 -48 V 42 0 V 0 -43 V 42 0 V 0 -209 V 42 0 V 0 81 V 42 0 V 0 -22 V 42 0 V 0 -73 V LT3 2394 1214 M 180 0 V 600 1437 M 42 0 V 0 50 V 42 0 V 0 -41 V 42 0 V 0 -46 V 42 0 V 0 -44 V 42 0 V 0 -46 V 42 0 V 0 -40 V 42 0 V 0 -35 V 42 0 V 0 -36 V 41 0 V 0 -28 V 42 0 V 0 -24 V 42 0 V 0 -32 V 42 0 V 0 -34 V 42 0 V 0 -14 V 42 0 V 0 -18 V 42 0 V 0 -31 V 42 0 V 0 -11 V 42 0 V 0 -36 V 42 0 V 0 -12 V 42 0 V 0 -27 V 42 0 V 0 -24 V 42 0 V 0 -27 V 42 0 V 0 2 V 42 0 V 0 -40 V 42 0 V 0 -2 V 42 0 V 0 -45 V 41 0 V 0 8 V 42 0 V 0 -15 V 42 0 V 0 -20 V 42 0 V 0 -12 V 42 0 V 0 -42 V 42 0 V 0 30 V 42 0 V 0 -42 V 42 0 V 0 -8 V 42 0 V 0 -7 V 42 0 V 0 -53 V 42 0 V 0 -12 V 42 0 V 0 36 V 42 0 V 0 -53 V 42 0 V 0 -3 V 42 0 V 0 7 V 42 0 V 0 -75 V 41 0 V 0 -12 V 42 0 V 0 69 V 42 0 V 42 0 V 0 -82 V 42 0 V 0 -93 V 42 0 V 0 63 V 42 0 V 0 -9 V 42 0 V 0 -54 V 42 0 V 0 100 V stroke grestore end showpage } \put(2334,1214){\makebox(0,0)[r]{Full DCM}} \put(2334,1314){\makebox(0,0)[r]{Parton shower}} \put(2334,1414){\makebox(0,0)[r]{LO+PS}} \put(1648,1677){\makebox(0,0){(b)}} \put(1648,-49){\makebox(0,0){$k_{\perp W}$ (GeV)}} \put(280,914){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$d\sigma/dk_{\perp W}$ (pb/GeV)}}% \special{ps: currentpoint grestore moveto}% } \put(2697,151){\makebox(0,0){100}} \put(2278,151){\makebox(0,0){80}} \put(1858,151){\makebox(0,0){60}} \put(1439,151){\makebox(0,0){40}} \put(1019,151){\makebox(0,0){20}} \put(600,151){\makebox(0,0){0}} \put(540,1577){\makebox(0,0)[r]{1000}} \put(540,1246){\makebox(0,0)[r]{100}} \put(540,914){\makebox(0,0)[r]{10}} \put(540,583){\makebox(0,0)[r]{1}} \put(540,251){\makebox(0,0)[r]{0.1}} \end{picture} \fcap{The transverse momentum spectrum of the $W$ at the Tevatron. In (a) the full line is the prediction of the full \tordas\ $W+$jet matrix element as implemented in \pythia. The dotted line is the prediction of the full DCM with $\beta=2$ and initial-state $\g2qq$\ splitting. In (b) the full line is as in (a) but with the parton shower of \pythia\ added after the first emission, the dashed line is \pythia\ using only parton showers and the dotted line is the same as in (a).} \label{figptW1} \end{figure} In fig.\refig{figptW1}a, the $W$ $\kt$\ spectrum at the Tevatron is shown, using the full \tordas\ matrix element (as implemented in \pythia) and using the modified DCM with the initial state $\g2qq$\ splitting as implemented in \ariadne\footnote{All results labelled \ariadne\ or DCM are actually generated using the zeroth-order $W$ production in \pythia, with the CDM added and, where indicated, using the string fragmentation implemented in \jetset\ \cite{JETPYT94}.}. Clearly the DCM does a good job of reproducing the high-$\kt$\ tail of the distribution. In fig.\refig{figptW1}b the DCM is compared with the two parton shower approaches of \pythia, one using only parton showers and one using first-order matrix elements with parton shower added. Since the DCM is a leading-log cascade, except that the first emission is uses the full matrix element, it smoothly interpolates between the pure parton shower description, which should be a good approximation for small $\kt$, and the matrix element description, which is good for high $\kt$, but has to be cut off at small $\kt$\ to avoid divergences. \section{Results and Predictions} \label{secres} In ref.\ \cite{GEOFFcdm94}, it was found that, when looking at the rapidity of the balancing jet in high-$\kt$\ $W$ events at the Tevatron, no correlation with the $W$ rapidity was found, while a leading order and a next to leading order calculation predicted a strong correlation. It was also found that a preliminary implementation of the DCM model described here reproduced data fairly well and only gave a very weak correlation\footnote{The results presented here differ from the ones in ref.\ \cite{GEOFFcdm94} due to a bug introduced in the initial-state $\g2qq$\ splitting in the preliminary version of \ariadne\ used in that paper.}. \begin{figure} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2880,1728)(-500,0) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 2097 0 V 600 251 M 0 1426 V LTb 600 251 M 63 0 V 2034 0 R -63 0 V 600 536 M 63 0 V 2034 0 R -63 0 V 600 821 M 63 0 V 2034 0 R -63 0 V 600 1107 M 63 0 V 2034 0 R -63 0 V 600 1392 M 63 0 V 2034 0 R -63 0 V 600 1677 M 63 0 V 2034 0 R -63 0 V 600 251 M 0 63 V 0 1363 R 0 -63 V 1019 251 M 0 63 V 0 1363 R 0 -63 V 1439 251 M 0 63 V 0 1363 R 0 -63 V 1858 251 M 0 63 V 0 1363 R 0 -63 V 2278 251 M 0 63 V 0 1363 R 0 -63 V 2697 251 M 0 63 V 0 1363 R 0 -63 V 600 251 M 2097 0 V 0 1426 V -2097 0 V 600 251 L LT0 1667 1534 M 180 0 V 810 415 M 419 355 V 420 315 V 419 216 V 419 97 V LT1 1667 1434 M 180 0 V 810 363 M 419 298 V 420 201 V 419 131 V 419 188 V LT3 1667 1334 M 180 0 V 810 369 M 419 152 V 420 144 V 419 129 V 419 -77 V stroke grestore end showpage } \put(1607,1334){\makebox(0,0)[r]{DCM}} \put(1607,1434){\makebox(0,0)[r]{LO+PS}} \put(1607,1534){\makebox(0,0)[r]{Leading order}} \put(1648,51){\makebox(0,0){$y_W$}} \put(280,964){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$\langle\eta_{\mbox{jet}}\rangle$}}% \special{ps: currentpoint grestore moveto}% } \put(2697,151){\makebox(0,0){2.5}} \put(2278,151){\makebox(0,0){2}} \put(1858,151){\makebox(0,0){1.5}} \put(1439,151){\makebox(0,0){1}} \put(1019,151){\makebox(0,0){0.5}} \put(600,151){\makebox(0,0){0}} \put(540,1677){\makebox(0,0)[r]{1}} \put(540,1392){\makebox(0,0)[r]{0.8}} \put(540,1107){\makebox(0,0)[r]{0.6}} \put(540,821){\makebox(0,0)[r]{0.4}} \put(540,536){\makebox(0,0)[r]{0.2}} \put(540,251){\makebox(0,0)[r]{0}} \end{picture} \fcap{The average jet pseudo-rapidity $\eta$ vs.\ the rapidity of the $W$ $y_W$ at the Tevatron. The jets are reconstructed with a cone algorithm using a radius of $0.7$, and in each event a jet with $E_\perp>20$ GeV, and $|\eta|<3$ on the oposite side in azimuth w.r.t.\ the $W$, is selected. In case of several such jets, the one with $\et$\ closest to the $\kt$\ of the $W$ is chosen. The full line is the leading order calculation as implemented in \pythia, the dashed line is the same, but with parton showers and fragmentation added, and the dotted line is the full DCM also with fragmentation added. (The ``kinkiness'' of the lines are due to limited statistics in the simulations.)} \label{figWjet1} \end{figure} Figure \ref{figWjet1} is an attempt to reconstruct the measurement in ref.\ \cite{GEOFFcdm94}\footnote{The details in the jet reconstruction may differ from that of ref.\ \cite{GEOFFcdm94}. In addition, the experimental ambiguity in the $y_W$ determination is not taken into account here.} on the generator level. As expected from eqs.\ (\ref{eqMEWg1}) and (\ref{eqMEWq1}), which are both symmetric around the $W$ rapidity, the leading-order calculation gives a more or less linear correlation between the mean jet pseudo-rapidity and the rapidity of the $W$, although $\langle\eta_{\mbox{\tiny jet}}\rangle\neq y_q$ due to the smearing of the structure function convolution and the limited kinematical acceptance for jets. When parton showers and fragmentation are added to the leading-order calculation, the smearing is increased. Also, since the phase space available for emissions is larger on the side where the $x$ of the incoming parton is smaller, the jets for large $y_W$ are ``dragged'' somewhat towards the centre, destroying the correlation. In the DCM, this dragging is more pronounced due to the ordering in the cascade as follows. \begin{figure} \setlength{\unitlength}{0.063mm} \begin{picture}(2400,700)(0,0) \put(50,100){\vector(1,0){1100}} \put(600,100){\vector(0,1){550}} \put(600,50){\makebox(0,0){(a)}} \put(100,100){\line(1,1){500}} \put(1100,100){\line(-1,1){500}} \put(750,350){\makebox(0,0){{\footnotesize $\bullet$}}} \put(950,450){\makebox(0,0){{\footnotesize $(y_W,\ln{m_W^2})$}}} \put(750,350){\vector(-1,-2){30}} \put(720,290){\vector(-1,-1){60}} \put(660,230){\vector(-2,-1){100}} \put(560,180){\vector(-4,-1){150}} \put(735,275){\makebox(0,0){{\footnotesize 1}}} \put(675,215){\makebox(0,0){{\footnotesize 2}}} \put(575,160){\makebox(0,0){{\footnotesize 3}}} \put(425,120){\makebox(0,0){{\footnotesize 4}}} \put(650,675){\makebox(0,0){{\footnotesize $\kappa$}}} \put(1175,70){\makebox(0,0){{\footnotesize $y$}}} \put(1250,100){\vector(1,0){1100}} \put(1800,100){\vector(0,1){550}} \put(1800,50){\makebox(0,0){(b)}} \put(1300,100){\line(1,1){500}} \put(2300,100){\line(-1,1){500}} \put(1300,100){\line(2,1){667}} \put(2300,100){\line(-2,1){667}} \put(1850,675){\makebox(0,0){{\footnotesize $\kappa$}}} \put(2375,70){\makebox(0,0){{\footnotesize $y$}}} \put(1950,350){\makebox(0,0){{\footnotesize $\bullet$}}} \put(2150,450){\makebox(0,0){{\footnotesize $(y_W,\ln{m_W^2})$}}} \put(1950,350){\vector(-1,-3){60}} \put(1890,150){\makebox(0,0){{\footnotesize 3}}} \put(1890,170){\vector(-1,1){100}} \put(1780,290){\makebox(0,0){{\footnotesize 1}}} \put(1790,270){\vector(-1,-1){120}} \put(1680,130){\makebox(0,0){{\footnotesize 4}}} \put(1670,150){\vector(-2,1){100}} \put(1550,200){\makebox(0,0){{\footnotesize 2}}} \end{picture} \fcap{Example of paths, tracing emissions backwards from the hard interaction $(y_W,\ln{m_W^2})$ on the small-$x$\ side of high-rapidity $W$ events for (a) the initial-state parton shower in \pythia, where the emissions are ordered both in $x$ and $\k2t$, and (b) the DCM, where the emissions, although the cascade is ordered in $\k2t$, when traced backwards from the hard interaction in this way, are ordered in $x$ but not in $\k2t$.} \label{figphase3} \end{figure} In the parton shower in \pythia, each step in the backward evolution of the initial-state shower is ordered in both $x$ and virtuality \tq2; thus even if the phase space is larger on the small-$x$\ side of the $W$, the shower quickly runs out of phase space due to the ordering in \tq2\ (resulting also in an ordering in $\k2t$) as in fig.\refig{figphase3}a. The DCM, although ordered in $\k2t$, is not ordered in $x$, or, if the final state partons are traced backwards in colour from the hard interaction, ordered in $x$ but {\em not} in $\k2t$ as in fig.\refig{figphase3}b. In this respect, the DCM is similar to the BFKL evolution, and it gives a good description of the large transverse energy flows in small-$x$\ events at HERA, which has been suggested as a signal for the BFKL evolution \cite{MARTINetflow94}. Because of this, the DCM can better use the increased phase space on the small-$x$\ side of the $W$ and the ``dragging'' effect is larger than for conventional parton showers in fig.\refig{figWjet1}, and the result closer to, if not consistent with, the measurement in ref.\ \cite{GEOFFcdm94}. The increase in transverse energy flow at small $x$ found at HERA should also be visible at the Tevatron in high-rapidity $W$ events. In fig.\refig{figEtflow1} the predictions for the $\et$\ flow from the parton shower model of \pythia\footnote{Since no high-$\kt$\ jets are required and the bulk of the events are at low $k_{\perp W}$, the matrix element plus parton shower approach in fig.\refig{figWjet1} is not adequate here.} and the DCM of \ariadne\ are shown for inclusive $W$ events at the Tevatron for two $W$ rapidity intervals. The two models are fairly similar at central $W$ rapidities, while for high $y_W$ the CDM gives more transverse energy, despite the fact that the hard interaction scale ($m_W^2\approx 6400$ $\mbox{GeV}^2$) is much larger here than at HERA ($\langle Q^2 \rangle \,\lower3pt\hbox{$\buildrel < \over\sim$}\, 100$ $\mbox{GeV}^2$). \begin{figure} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2160,1728)(100,0) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 1377 0 V -688 0 R 0 1326 V LTb 600 251 M 63 0 V 1314 0 R -63 0 V 600 781 M 63 0 V 1314 0 R -63 0 V 600 1312 M 63 0 V 1314 0 R -63 0 V 600 251 M 0 63 V 0 1263 R 0 -63 V 830 251 M 0 63 V 0 1263 R 0 -63 V 1059 251 M 0 63 V 0 1263 R 0 -63 V 1289 251 M 0 63 V 0 1263 R 0 -63 V 1518 251 M 0 63 V 0 1263 R 0 -63 V 1748 251 M 0 63 V 0 1263 R 0 -63 V 1977 251 M 0 63 V 0 1263 R 0 -63 V 600 251 M 1377 0 V 0 1326 V -1377 0 V 600 251 L LT0 1693 649 M 180 0 V 657 745 M 772 852 L 887 980 L 115 143 V 114 73 V 115 68 V 115 30 V 115 -45 V 114 -100 V 115 -119 V 1805 887 L 1920 778 L LT3 1693 549 M 180 0 V 657 827 M 115 78 V 115 88 V 115 94 V 114 99 V 115 93 V 115 -32 V 115 -50 V 114 -77 V 1690 995 L 115 -91 V 115 -75 V stroke grestore end showpage } \put(1633,549){\makebox(0,0)[r]{DCM}} \put(1633,649){\makebox(0,0)[r]{PS}} \put(1288,1677){\makebox(0,0){(a) $0.0<y_W<0.5$}} \put(1288,51){\makebox(0,0){$\eta$}} \put(400,914){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$dE_\perp/d\eta$ (GeV)}}% \special{ps: currentpoint grestore moveto}% } \put(1977,151){\makebox(0,0){3}} \put(1748,151){\makebox(0,0){2}} \put(1518,151){\makebox(0,0){1}} \put(1289,151){\makebox(0,0){0}} \put(1059,151){\makebox(0,0){-1}} \put(830,151){\makebox(0,0){-2}} \put(600,151){\makebox(0,0){-3}} \put(540,1312){\makebox(0,0)[r]{4}} \put(540,781){\makebox(0,0)[r]{2}} \put(540,251){\makebox(0,0)[r]{0}} \end{picture} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2160,1728)(100,0) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 1377 0 V -688 0 R 0 1326 V LTb 600 251 M 63 0 V 1314 0 R -63 0 V 600 781 M 63 0 V 1314 0 R -63 0 V 600 1312 M 63 0 V 1314 0 R -63 0 V 600 251 M 0 63 V 0 1263 R 0 -63 V 830 251 M 0 63 V 0 1263 R 0 -63 V 1059 251 M 0 63 V 0 1263 R 0 -63 V 1289 251 M 0 63 V 0 1263 R 0 -63 V 1518 251 M 0 63 V 0 1263 R 0 -63 V 1748 251 M 0 63 V 0 1263 R 0 -63 V 1977 251 M 0 63 V 0 1263 R 0 -63 V 600 251 M 1377 0 V 0 1326 V -1377 0 V 600 251 L LT0 1693 649 M 180 0 V 657 613 M 115 33 V 115 46 V 115 58 V 114 105 V 115 40 V 115 46 V 115 51 V 1575 890 L 115 -92 V 115 -58 V 115 -47 V LT3 1693 549 M 180 0 V 657 845 M 115 50 V 115 61 V 115 60 V 114 67 V 115 48 V 115 22 V 115 19 V 114 -95 V 115 -60 V 1805 888 L 1920 784 L stroke grestore end showpage } \put(1633,549){\makebox(0,0)[r]{DCM}} \put(1633,649){\makebox(0,0)[r]{PS}} \put(1288,1677){\makebox(0,0){(b) $2.0<y_W<2.5$}} \put(1288,51){\makebox(0,0){$\eta$}} \put(400,914){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$dE_\perp/d\eta$ (GeV)}}% \special{ps: currentpoint grestore moveto}% } \put(1977,151){\makebox(0,0){3}} \put(1748,151){\makebox(0,0){2}} \put(1518,151){\makebox(0,0){1}} \put(1289,151){\makebox(0,0){0}} \put(1059,151){\makebox(0,0){-1}} \put(830,151){\makebox(0,0){-2}} \put(600,151){\makebox(0,0){-3}} \put(540,1312){\makebox(0,0)[r]{4}} \put(540,781){\makebox(0,0)[r]{2}} \put(540,251){\makebox(0,0)[r]{0}} \end{picture} \fcap{The transverse energy flow in inclusive $W$ events at the Tevatron for (a) $0.0<y_W<0.5$ and (b) $2.0<y_W<2.5$. The full and dotted lines are the predictions of the DCM in \ariadne\ and of the parton shower in \pythia, respectively.} \label{figEtflow1} \end{figure} In $pp$ collisions we also have to worry about underlying events. In fig.\refig{figWjet1}, this does not give large effects since a large-$\et$\ jet is required, but for fig.\refig{figEtflow1} the underlying event would give an extra contribution to the $\et$\ flow. This extra contribution should however be evenly spread out in $\eta$ and independent of $y_W$, and the differences between the parton shower and DCM approaches should survive. To take this contribution into account, the multiple interaction model implemented in \pythia\ \cite{TSmult87} has been used. Note, however, that in the case of the DCM, only the qualitative features of the contribution are completely relevant, as the parameters of the multiple interaction model probably need to be retuned to fit the DCM. \begin{figure} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2160,1728)(100,0) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 1377 0 V 600 251 M 0 1326 V LTb 600 251 M 63 0 V 1314 0 R -63 0 V 600 630 M 63 0 V 1314 0 R -63 0 V 600 1009 M 63 0 V 1314 0 R -63 0 V 600 1388 M 63 0 V 1314 0 R -63 0 V 600 251 M 0 63 V 0 1263 R 0 -63 V 875 251 M 0 63 V 0 1263 R 0 -63 V 1151 251 M 0 63 V 0 1263 R 0 -63 V 1426 251 M 0 63 V 0 1263 R 0 -63 V 1702 251 M 0 63 V 0 1263 R 0 -63 V 1977 251 M 0 63 V 0 1263 R 0 -63 V 600 251 M 1377 0 V 0 1326 V -1377 0 V 600 251 L LT0 1486 1198 M 180 0 V 738 680 M 275 46 V 276 27 V 275 -3 V 275 -39 V LT3 1486 1098 M 180 0 V 738 718 M 275 65 V 276 44 V 275 33 V 275 20 V stroke grestore end showpage } \put(1426,1098){\makebox(0,0)[r]{DCM}} \put(1426,1198){\makebox(0,0)[r]{PS}} \put(1288,1677){\makebox(0,0){(a)}} \put(1288,51){\makebox(0,0){$y_W$}} \put(400,914){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$dE_\perp/d\eta$ (GeV)}}% \special{ps: currentpoint grestore moveto}% } \put(1977,151){\makebox(0,0){2.5}} \put(1702,151){\makebox(0,0){2}} \put(1426,151){\makebox(0,0){1.5}} \put(1151,151){\makebox(0,0){1}} \put(875,151){\makebox(0,0){0.5}} \put(600,151){\makebox(0,0){0}} \put(540,1388){\makebox(0,0)[r]{6}} \put(540,1009){\makebox(0,0)[r]{4}} \put(540,630){\makebox(0,0)[r]{2}} \put(540,251){\makebox(0,0)[r]{0}} \end{picture} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2160,1728)(100,0) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 1377 0 V 600 251 M 0 1326 V LTb 600 251 M 63 0 V 1314 0 R -63 0 V 600 630 M 63 0 V 1314 0 R -63 0 V 600 1009 M 63 0 V 1314 0 R -63 0 V 600 1388 M 63 0 V 1314 0 R -63 0 V 600 251 M 0 63 V 0 1263 R 0 -63 V 875 251 M 0 63 V 0 1263 R 0 -63 V 1151 251 M 0 63 V 0 1263 R 0 -63 V 1426 251 M 0 63 V 0 1263 R 0 -63 V 1702 251 M 0 63 V 0 1263 R 0 -63 V 1977 251 M 0 63 V 0 1263 R 0 -63 V 600 251 M 1377 0 V 0 1326 V -1377 0 V 600 251 L LT0 1486 630 M 180 0 V 738 1094 M 275 50 V 276 18 V 275 -15 V 275 -81 V LT3 1486 530 M 180 0 V 738 1213 M 275 73 V 276 46 V 275 66 V 275 -44 V stroke grestore end showpage } \put(1426,530){\makebox(0,0)[r]{DCM+MI}} \put(1426,630){\makebox(0,0)[r]{PS+MI}} \put(1288,1677){\makebox(0,0){(b)}} \put(1288,51){\makebox(0,0){$y_W$}} \put(400,914){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$dE_\perp/d\eta$ (GeV)}}% \special{ps: currentpoint grestore moveto}% } \put(1977,151){\makebox(0,0){2.5}} \put(1702,151){\makebox(0,0){2}} \put(1426,151){\makebox(0,0){1.5}} \put(1151,151){\makebox(0,0){1}} \put(875,151){\makebox(0,0){0.5}} \put(600,151){\makebox(0,0){0}} \put(540,1388){\makebox(0,0)[r]{6}} \put(540,1009){\makebox(0,0)[r]{4}} \put(540,630){\makebox(0,0)[r]{2}} \put(540,251){\makebox(0,0)[r]{0}} \end{picture} \fcap{The transverse energy flow two units of rapidity ``behind'' the $W$ for inclusive $W$ events, at the Tevatron, i.e.\ for the interval $1.0<y_W<1.5$ the $\et$\ flow in the pseudo-rapidity interval $-1.5<\eta<-1.0$. The full line is the \pythia\ parton shower and the dotted line is the DCM without (a) and with (b) multiple interactions.} \label{figEtflow2} \end{figure} The differences between the parton shower and the DCM approaches are most significant when the $\et$\ flow is measured as a function of $y_W$ as in fig.\refig{figEtflow2}, where the $\et$\ flow, two units of rapidity away from the $W$, is shown for both models, with and without the multiple interaction model for the underlying event implemented in \pythia\ \cite{TSmult87}. It is clear that the underlying event introduces an extra $\et$\ for both models, but that the dependence of the $\et$\ flow on $y_W$ is still different for the two models. As expected the $\et$\ flow is more or less constant for the parton shower approach, but increases slightly for the DCM due to the increase in phase space at large $y_W$. \begin{figure} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2160,1728)(100,0) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 1377 0 V 600 251 M 0 1326 V LTb 600 251 M 63 0 V 1314 0 R -63 0 V 600 630 M 63 0 V 1314 0 R -63 0 V 600 1009 M 63 0 V 1314 0 R -63 0 V 600 1388 M 63 0 V 1314 0 R -63 0 V 600 251 M 0 63 V 0 1263 R 0 -63 V 875 251 M 0 63 V 0 1263 R 0 -63 V 1151 251 M 0 63 V 0 1263 R 0 -63 V 1426 251 M 0 63 V 0 1263 R 0 -63 V 1702 251 M 0 63 V 0 1263 R 0 -63 V 1977 251 M 0 63 V 0 1263 R 0 -63 V 600 251 M 1377 0 V 0 1326 V -1377 0 V 600 251 L LT0 1486 1198 M 180 0 V 738 680 M 275 -30 V 276 -39 V 275 -38 V 275 -40 V LT3 1486 1098 M 180 0 V 738 647 M 275 -5 V 276 1 V 275 -3 V 275 -3 V stroke grestore end showpage } \put(1426,1098){\makebox(0,0)[r]{DCM}} \put(1426,1198){\makebox(0,0)[r]{PS}} \put(1288,1677){\makebox(0,0){(a)}} \put(1288,51){\makebox(0,0){$y_W$}} \put(400,914){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$dE_\perp/d\eta$ (GeV)}}% \special{ps: currentpoint grestore moveto}% } \put(1977,151){\makebox(0,0){2.5}} \put(1702,151){\makebox(0,0){2}} \put(1426,151){\makebox(0,0){1.5}} \put(1151,151){\makebox(0,0){1}} \put(875,151){\makebox(0,0){0.5}} \put(600,151){\makebox(0,0){0}} \put(540,1388){\makebox(0,0)[r]{6}} \put(540,1009){\makebox(0,0)[r]{4}} \put(540,630){\makebox(0,0)[r]{2}} \put(540,251){\makebox(0,0)[r]{0}} \end{picture} \setlength{\unitlength}{0.1bp} \special{! /gnudict 40 dict def gnudict begin /Color false def /Solid false def /gnulinewidth 5.000 def /vshift -33 def /dl {10 mul} def /hpt 31.5 def /vpt 31.5 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 /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 /PL { stroke gnulinewidth 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 /P { stroke [] 0 setdash currentlinewidth 2 div sub M 0 currentlinewidth V stroke } def /D { 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 P } def /A { stroke [] 0 setdash vpt sub M 0 vpt2 V currentpoint stroke M hpt neg vpt neg R hpt2 0 V stroke } def /B { 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 P } def /C { 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 /T { 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 P } def /S { 2 copy A C} def end } \begin{picture}(2160,1728)(100,0) \special{" gnudict begin gsave 50 50 translate 0.100 0.100 scale 0 setgray /Helvetica findfont 100 scalefont setfont newpath -500.000000 -500.000000 translate LTa 600 251 M 1377 0 V 600 251 M 0 1326 V LTb 600 251 M 63 0 V 1314 0 R -63 0 V 600 630 M 63 0 V 1314 0 R -63 0 V 600 1009 M 63 0 V 1314 0 R -63 0 V 600 1388 M 63 0 V 1314 0 R -63 0 V 600 251 M 0 63 V 0 1263 R 0 -63 V 875 251 M 0 63 V 0 1263 R 0 -63 V 1151 251 M 0 63 V 0 1263 R 0 -63 V 1426 251 M 0 63 V 0 1263 R 0 -63 V 1702 251 M 0 63 V 0 1263 R 0 -63 V 1977 251 M 0 63 V 0 1263 R 0 -63 V 600 251 M 1377 0 V 0 1326 V -1377 0 V 600 251 L LT0 1486 725 M 180 0 V 738 1094 M 275 -44 V 276 -40 V 275 -56 V 275 -79 V LT3 1486 625 M 180 0 V 738 1213 M 275 -3 V 276 -12 V 275 -5 V 275 -52 V LT1 1486 525 M 180 0 V 738 1391 M 275 -58 V 276 -87 V 275 -57 V 275 -45 V LT4 1486 425 M 180 0 V 738 1138 M 275 5 V 276 5 V 275 -3 V 275 -34 V stroke grestore end showpage } \put(1426,425){\makebox(0,0)[r]{DCM+MI $\beta=0$}} \put(1426,525){\makebox(0,0)[r]{ME+PS+MI}} \put(1426,625){\makebox(0,0)[r]{DCM+MI}} \put(1426,725){\makebox(0,0)[r]{PS+MI}} \put(1288,1677){\makebox(0,0){(b)}} \put(1288,51){\makebox(0,0){$y_W$}} \put(400,914){% \special{ps: gsave currentpoint currentpoint translate 270 rotate neg exch neg exch translate}% \makebox(0,0)[b]{\shortstack{$dE_\perp/d\eta$ (GeV)}}% \special{ps: currentpoint grestore moveto}% } \put(1977,151){\makebox(0,0){2.5}} \put(1702,151){\makebox(0,0){2}} \put(1426,151){\makebox(0,0){1.5}} \put(1151,151){\makebox(0,0){1}} \put(875,151){\makebox(0,0){0.5}} \put(600,151){\makebox(0,0){0}} \put(540,1388){\makebox(0,0)[r]{6}} \put(540,1009){\makebox(0,0)[r]{4}} \put(540,630){\makebox(0,0)[r]{2}} \put(540,251){\makebox(0,0)[r]{0}} \end{picture} \fcap{The transverse energy flow in the interval $-2.5<\eta<-2.0$ for $W$ events at the Tevatron as a function of $y_W$. The full line is the \pythia\ parton shower and the dotted line is the DCM without (a) and with (b) multiple interactions. In (b) the dashed line is \pythia\ using \tordas\ matrix elements with addition of parton showers and multiple interactions, and the dash-dotted line is the DCM with $\beta=\infty$.} \label{figEtflow3} \end{figure} Similarly, looking at the $\et$\ flow in a fixed rapidity interval for varying $y_W$, the DCM predicts a fairly constant value, while in the parton shower approach, the flow decreases with increasing $y_W$ as in fig.\refig{figEtflow3}. To check that the differences are not due to the fact that the DCM has the correct \tordas\ matrix element in the first emission, fig.\refig{figEtflow3}b also contains a line with the matrix element plus parton shower option in \pythia. It is clear that, although the $\et$\ flow is higher since only events with $k_{\perp W}>10$ GeV are included, it has the same $y_W$ dependence as the plain parton shower approach. Also in fig.\refig{figEtflow3}, the changing of the $\beta$ parameter in the DCM is shown to have some effect on the $\et$\ flow; however, the dependence on $y_W$ is still the same. \section{Conclusions} \label{secsum} The model presented in this paper is not perfect. The treatment of initial-state $\g2qq$\ splitting is a bit foreign to the original Colour Dipole Model, and so is the transfer of transverse recoil to the $W$ for gluon emissions. But as a leading log approximation it should be just as good as the conventional parton shower approach, and it has the advantage of correctly describing also high-$\kt$\ $W$ production. In addition it gives the opportunity of studying effects of unordered parton evolution on the hadronic final state, and, although there are some uncertainties in the overall $\et$\ level due to multiple interactions and the $\beta$ parameter in the DCM, it would be very interesting to compare the predictions for the $y_W$ dependence of the $\et$\ flow presented in this paper with data from the Tevatron. Although only $W$ production at the Tevatron has been discussed in this paper, the model can of course be applied to any Drell--Yan-like process in any hadron--hadron collision. And, as pointed out above, the initial-state $\g2qq$\ splitting can be used also in deep inelastic lepton--hadron scattering. The model presented here is also the ``last piece'' to complete the DCM description of QCD cascades for all standard processes in $\ee$, $ep$ and $pp$ collisions. This is reflected in the fact that \ariadne\ now is fully interfaced to all hard sub-processes in \pythia. However, some care must be taken when using the multiple interaction model of \pythia\ together with the DCM, as mentioned above. \section*{Acknowledgements} I would like to thank Bo Andersson and G\"{o}sta Gustafson for valuable discussions.
\section{Introduction} \label{introduction} The study of the phase diagram of fundamental--adjoint pure gauge systems, \begin{equation} S = \beta_f \sum_{P} [1 - \frac{1}{N} {\rm Re} {\rm Tr} U_{P} ] + \beta_a \sum_{P} [1 - \frac{1}{N^2} {\rm Tr} U^\dagger_{P} {\rm Tr} U_{P}] , \label{eq:aaction} \end{equation} in the early 80's revealed a non-trivial phase structure with first order (bulk) transitions in the region of small $\beta_f$ \cite{Greensite81,Bhanot82}. In particular a line of first order transitions points towards the fundamental axis and, after terminating, extends as a roughly straight line of bulk crossovers to the fundamental axis and beyond. This non-trivial phase structure, and in particular the critical endpoint, has been argued to be associated with the dip in the discrete $\beta$-function of the theory with standard Wilson action, which occurs in the region where the bulk crossover line crosses the fundamental axis. In a couple of recent papers Gavai, Grady and Mathur \cite{Gavai94} and Mathur and Gavai \cite{Gavai94b} returned to investigating the behavior of pure gauge SU(2) theory in the fundamental--adjoint plane at finite, non-zero temperature. They raised doubts about the bulk nature of the phase transition and claimed that their results were consistent with the transitions, for lattices with temporal extent $N_t = 4$, 6 and 8, to be of thermal, deconfining nature, displaced toward weak coupling with increasing $N_t$. On the transition line for a fixed $N_t$ there is then a switch from second order behavior near the fundamental axis to first order behavior at larger adjoint coupling. In a Landau Ginzburg model of the effective action in terms of Polyakov lines, Mathur \cite{Mathur95} reported that he could reproduce the claimed behavior seen in the numerical simulations. These results, should they be confirmed, are rather unsettling, since they contradict the usual universality picture of lattice gauge theory with a second order deconfinement transition for gauge group SU(2). Puzzled by the finding of Ref.~\cite{Gavai94}, we studied the finite temperature behavior of pure gauge SU(3) theory in the fundamental--adjoint plane \cite{SU3_F_A,SU3_F_Ab}. We obtained results in agreement with the usual universality picture: there is a first order bulk transition line ending at \begin{equation} (\beta^*_f,\beta^*_a) = (4.00(7), 2.06(8)) . \label{eq:endpoint} \end{equation} The thermal deconfinement transition lines for fixed $N_t$ (being of first order for gauge group SU(3)) in the fundamental adjoint plane are ordered such that the thermal transition for smaller $N_t$ occurs to the left, at smaller $\beta_f$, than that for a larger $N_t$. In this order the thermal transition lines join on to the bulk transition line. The thermal transition line for $N_t = 4$ joins the bulk transition line very close to the critical endpoint. This is shown in Figure 4 of Ref.~\cite{SU3_F_Ab}, reproduced here as Figure~\ref{fig:phas_diag_T}. \begin{figure} \begin{center} \vskip 30mm \leavevmode \epsfysize=360pt \epsfbox[90 40 580 490]{phas_diag_T.eps} \end{center} \caption{The phase diagram together with the thermal deconfinement transition points for $N_t=2$, 4, 6 and 8 from Ref.~\protect\cite{SU3_F_Ab}. The lower plot shows an enlargement of the region around the end point of the bulk transition.} \label{fig:phas_diag_T} \end{figure} To solidify this picture, which is in agreement with the usual universality scenario, we have continued the investigation studying zero-temperature observables. We computed the string tension and the masses of some glueballs, in particular the $0^{++}$ glueball, along the thermal transition line for $N_t = 4$. For universal continuum behavior $\sqrt{\sigma}$ and the glueball masses should be constant along the thermal transition line for a fixed $N_t$, leading to constant ratios $T_c/\sqrt{\sigma}$ and $m_g/\sqrt{\sigma}$. Of course, at small $N_t$, corresponding to a large lattice spacing $a$, we expect to see some deviations from this constant behavior. However, we find large deviations for $m_{0^{++}}/\sqrt{\sigma}$ as we approach the critical endpoint of the bulk transition along the $N_t=4$ thermal transition line. The scalar glueball mass decreases dramatically, much more than what could be expected from simple scaling violations at a large lattice spacing. On the other hand, this is not really a surprise since at the critical endpoint at least one mass in the $0^{++}$-channel has to vanish. We elaborate on our findings in the next sections and then discuss the implications for the scaling behavior along the fundamental (or Wilson) axis. \section{Observables and Analysis} \label{analysis} We have made simulations of the model with action (\ref{eq:aaction}) along the thermal transition line for $N_t=4$, and continued along the bulk transition line, on a $12^4$ lattice. Experience indicates this size to be sufficient to avoid significant finite size effects. The simulations were carried out with a 10-hit Metropolis algorithm tuned to an acceptance rate of about 50\%. Observables were measured every 20 sweeps after thermalization. Close to the fundamental axis this resulted in essentially statistically independent measurements, while closer to the critical endpoint the autocorrelation time was significantly increased. We computed finite $T$ approximants to the potential from time-like Wilson loops constructed with `APE'-smeared spatial links \cite{APE_smear} to increase the overlap to the ground state potential. On and off axis spatial paths were considered with distances $R = n$, $\sqrt{2} n$, $\sqrt{3} n$ and $\sqrt{5} n$, with $n = 1$, 2, \dots an integer. The string tension was then extracted from the usual fit \begin{equation} V(\vec R) = V_0 - \frac{e}{R} + l \left( G_L(\vec R) - \frac{1}{R} \right) + \sigma R . \label{eq:fit_form} \end{equation} Here $G_L(\vec R)$ is the lattice Coulomb potential, included in the fit to take account of short distance lattice artefacts. Our fits are fully correlated $\chi^2$-fits with the correlations estimated by bootstrap, after binning to alleviate autocorrelation effects. The results of the best fits are listed in Table~\ref{tab:fits}. \begin{table}[ht] \centerline{% \begin{tabular}{|l|l|l|l|r|l|l|c|} \hline $\beta_a$ & $\beta_{fc}(N_t=4)$ & $\beta_f$ & $L$ & $N_{meas}$ & $\sqrt{\sigma}$ & $m_{0^{++}}$ & $m_{2^{++}}(t=1)$ \\ \hline 0.0 & 5.6925(2) & 5.7$^a$ & & & 0.4099( 12) & 0.97( 2) & 2.39(13) \\ 0.5 & 5.25(5) & 5.25 & 12 & 500 & 0.4218( 28) & 0.93(11) & 2.29(13) \\ 1.0 & 4.85(5) & 4.85 & 12 & 500 & 0.4024( 82) & 0.78(28) & 2.45(19) \\ 1.5 & 4.45(5) & 4.45 & 12 & 500 & 0.3743( 51) & 0.56(17) & 2.13( 9) \\ 2.0 & 4.035(5) & 4.03 & 12 & 1000 & 0.555 ( 11) & 0.34( 6) & 3.11(16) \\ 2.0 & 4.035(5) & 4.035 & 12 & 2000 & 0.4725(128) & 0.20( 4) & 3.17(14) \\ 2.0 & 4.035(5) & 4.04 & 12 & 1000 & 0.3750( 24) & 0.27( 8) & 2.37( 9) \\ \hline 2.25 & 3.8475(25) & 3.8475$^b$ & 12 & 500 & 0.619 ( 18) & 0.37(10) & 3.15(37) \\ 2.25 & 3.8475(25) & 3.8475$^c$ & 12 & 500 & 0.3005( 22) & 0.61( 8) & 1.59( 7) \\ 2.25 & 3.8475(25) & 3.8475$^c$ & 16 & 500 & 0.2965( 19) & 0.62( 5) & 1.72( 9) \\ \hline \end{tabular}% \caption{The results in the neighborhood of the $N_t=4$ thermal transition line. Comments: (a) $\protect\sqrt{\sigma}$ at $\beta_a=0.0$ comes from \protect\cite{MTc_sig}, $m_{0^{++}}$ from \protect\cite{GF11_gb}; we did not take their best value, but rather the effective mass from the same distance as at $\beta_a=0.5$; $m_{2^{++}}$ is from \protect\cite{dFSST86}; (b) in the disordered phase and (c) in the ordered phase on the bulk transition.} \label{tab:fits} \medskip\noindent \end{table} We also computed glueball correlation functions in the $0^{++}$, $2^{++}$ and $1^{+-}$ channel that can be built from simple plaquette operators. In an attempt to improve the signals we built these plaquettes not only from the original, but also from the smeared links, already used for the computation of the potential. Not surprisingly for computations done around the critical coupling for the $N_t = 4$ thermal phase transition, we did not obtain a significant signal in the $1^{+-}$ channel and only an effective mass from time distances $t = 0/1$ in the $2^{++}$ channel -- we had 500 measurements everywhere, except for $\beta_a = 2.0$, near the critical endpoint where the number was increased, as given in Table~\ref{tab:fits}. In the $0^{++}$ channel we got a signal at distance $t = 1/2$ at small $\beta_a$ and out to $t = 3/4$ at $\beta_a = 2.0$. Our best results are also given in Table~\ref{tab:fits}. The quantities $\sqrt{\sigma}$ and $m_{0^{++}}$ are shown in Figure~\ref{fig:m_and_sig} plotted versus $\beta_a$. As can be seen, $\sqrt{\sigma}$ remains approximately constant along the thermal transition line for $N_t=4$ -- the errors shown in the figure are statistical only; no estimate of the error from the uncertainty in the determination of $\beta_{fc}$ has been attempted except for $\beta_a=2.0$. There, the computation has been repeated for two nearby couplings, also listed in Table~\ref{tab:fits}; the variation with $\beta_f$ becomes so rapid that the error in the determination of $\beta_{fc}$ becomes the dominating factor. While our estimate for $m_{2^{++}}$, albeit an unreliable estimate since we had to use distances $t=0$ and 1 to obtain enough of a signal to extract an effective mass, also remains approximately constant, $m_{0^{++}}$ shows a remarkable decrease as the critical endpoint is approached. This observed behavior suggests that the mass in the $0^{++}$ channel vanishes at the critical endpoint, thereby giving strong additional evidence for the existence of this critical endpoint, since at a critical point at least one mass, in the $0^{++}$ channel to have a rotationally invariant continuum limit, must vanish. Since no other observable seems to be dramatically affected by the critical endpoint, we conjecture that the continuum theory one would obtain there is simply the (trivial) $\phi^4$ theory. To substantiate this claim somewhat, we made a fit to the scalar mass of the form \begin{equation} m_{0^{++}} = A \left( \beta^*_a - \beta_a \right)^p \label{eq:massfit} \end{equation} expected near a critical point. A 3-parameter fit gave $A=0.76(11)$, $p=0.35(20)$ and $\beta^*_a=2.02(6)$ with a $\chi^2$ of $0.29$ for 2 dof. Note that the estimate for $\beta^*_a$ is in agreement with the previous estimate (\ref{eq:endpoint}) obtained in \cite{SU3_F_Ab} from fits to the jump in the plaquette across the bulk transition line. Within its large error, the exponent $p$ is compatible with the mean field value $0.5$ of $\phi^4$ theory, up to logarithmic corrections. Since the errors of the fit parameters are rather large we also made a fit with $\beta^*_a$ held fixed at its value $2.06$ of (\ref{eq:endpoint}). This fit gave $A=0.71(3)$ and $p=0.44(5)$ with a $\chi^2$ of $0.47$ for 3 dof. Again, $p$ is compatible with the mean field value. Finally, a fit with $\beta^*_a=2.06$ and $p=0.5$ both held fixed gave $A=0.68(1)$ with a $\chi^2$ of $1.50$ for 4 dof. This last, still very acceptable fit is included in Figure~\ref{fig:m_and_sig}. \begin{figure} \begin{center} \leavevmode \epsfysize=360pt \epsfbox[90 40 580 490]{m_and_sig.eps} \end{center} \caption{$\protect\sqrt{\sigma}$ (octagons) and $m_{0^{++}}$ (squares) as a function of $\beta_a$ along the thermal transition line for $N_t=4$. At $\beta_a=2.25$ we show the results from both phases at the bulk transition. The dashed vertical line gives the location of the critical endpoint, (\protect\ref{eq:endpoint}), with the dotted lines indicating the error band. The curve is a fit to $m_{0^{++}}=A(2.06-\beta_a)^{1/2}$.} \label{fig:m_and_sig} \end{figure} \begin{figure} \begin{center} \leavevmode \epsfysize=360pt \epsfbox[90 40 580 490]{pot_b3p8475_ba2p25.eps} \end{center} \caption{The potential in both phases on the bulk transition line at $(\beta_f,\beta_a)=(3.8475,2.25)$ on $12^4$ lattices ($\Box$ and $\times$) and on a $16^4$ lattice ($\circ$).} \label{fig:pot_ba2p25} \end{figure} While we believe to have assembled impressive further evidence for the bulk nature of the phase transition line in the fundamental--adjoint plane and of its critical endpoint, we decided to make one further test. We computed the potential in both phases on the bulk transition line at $\beta_a = 2.25$. This is about the location where the $N_t=6$ thermal transition line joins onto the bulk transition line according to Ref.~\cite{SU3_F_Ab}. Hence, even in the ordered phase a lattice of size $12^4$ should be large enough to obtain a reliable potential. To make sure that this is indeed the case, we repeated the computation in the ordered phase on a $16^4$ lattice. As can be seen in Table~\ref{tab:fits} the finite size effect on the $12^4$ lattice does indeed appear negligible. On both sides of the bulk transition -- at identical couplings: we did not observe even an attempt at tunneling from one phase to the other -- we find a clearly confining potential. Therefore the bulk transition does not, as expected, affect confinement, provided the lattices are large enough. Of course, the string tension (and glueball masses) jump from one side of the bulk transition to the other. Indeed, as the bulk transition line is approached from the fundamental axis the string tension varies more and more rapidly -- and the thermal deconfinement transition lines for different $N_t$ come closer together -- {\it i.e.,} the dip in the step $\beta$-function becomes deeper until a jump is developed through the bulk transition line. \section{Implications for Scaling} \label{scaling} In the previous section we have corroborated the existence of a first order bulk phase transition line ending in a critical endpoint. We have provided evidence that physical observables are little affected by this phase transition line, except for a stronger dependence on $\beta_f$ at fixed $\beta_a$ -- a deepening of the dip in the step $\beta$-function -- and eventually a jump across the bulk transition line. The notable exception to this is the glueball mass in the $0^{++}$ channel which decreases as the critical endpoint is approached and vanishes there. \begin{figure} \begin{center} \leavevmode \epsfysize=360pt \epsfbox[90 40 580 490]{m0_ov_sig.eps} \end{center} \caption{$10T_c/\protect\sqrt{\sigma}$ (octagons) and $m_{0^{++}}/\protect\sqrt{\sigma}$ (squares) as a function of $\beta$ for the fundamental (Wilson) action.} \label{fig:m0_ov_sig} \end{figure} The influence of the critical endpoint on the $0^{++}$ glueball appears still visible in the crossover region on the fundamental (Wilson) axis. This has first been argued in Ref.~\cite{Rajan86}. The $0^{++}$ glueball is lighter in the crossover region, leading to a different scaling behavior than other observables. This can be seen in Figure~\ref{fig:m0_ov_sig} where we show the latest data of $T_c/\sqrt{\sigma}$ from Ref.~\cite{Boyd95} and $m_{0^{++}}/\sqrt{\sigma}$ with $\sqrt{\sigma}$ taken from Refs.~\cite{MTc_sig,B-S_sig} and the glueball mass from Refs.~\cite{GF11_gb,M-T_gb,Bali_gb}. While $T_c/\sqrt{\sigma}$ stays approximately constant in the $\beta$ interval shown, $m_{0^{++}}/\sqrt{\sigma}$ decreases visibly in the crossover region around $\beta = 5.7$. It has long been known that the scalar glueball mass scales differently in the crossover region than $T_c$ and the string tension, whose scaling behavior as a function of $\beta$ deviates from asymptotic scaling but agrees with the step $\beta$-function found in MCRG computations. The scalar glueball seemed more compatible with asymptotic scaling. However, in view of our findings that the different behavior of the scalar glueball mass comes from the influence of the nearby critical endpoint this asymptotic scaling behavior appears to be accidental. It would be interesting to determine the scaling behavior of other (pure gauge) observables. We conjecture that they would behave more like $T_c$ or $\sqrt{\sigma}$ than like the scalar glueball mass. Unfortunately no reliable, large volume, glueball masses for $J^{PC}$ quantum numbers other than $0^{++}$ for couplings in the crossover region exist in the literature to check this conjecture. \section*{Acknowledgements} This work was partly supported by the DOE under grants \#~DE-FG05-85ER250000 and \#~DE-FG05-92ER40742. The computations were carried out on the cluster of IBM RS6000's at SCRI.
\section{The Gauged Linear Sigma Model} We begin with a brief review of the gauged linear sigma model (GLSM) construction of Witten\cite{phases}. The model is formulated in (2,2) superspace, and requires for its construction the specification of a compact abelian group $G$, a faithful representation $\rho:G\toU(1)^n$, and a $G$-invariant polynomial $W(x_1,\dots,x_n)$, where $x_1, \dots, x_n$ are coordinates in a complex vector space ${\bf} \def\bf{\bf C}^n$ on which $U(1)^n$ acts diagonally. To construct the gauged linear sigma model, we begin with $n$ chiral superfields $\Phi_i$ (satisfying $\overline{D}_+\Phi_i = \overline{D}_-\Phi_i=0$) interacting via the holomorphic superpotential $W(\Phi_1,\ldots,\Phi_n)$. The model is invariant under the action of $G$ (via $\rho$) on $\vec\Phi$ and we gauge this action, preserving $N{=}2$ supersymmetry, by introducing the ${\bf g}$-valued vector multiplet $V$ with invariant field strength $\Sigma=\frac1{\sqrt{2}}\overline{D}_+D_-V$. This last field is {\it twisted chiral}, which means that $\overline{D}_+\Sigma =D_-\Sigma=0$. For each continuous $U(1)$ factor of $G$ we include a Fayet-Illiopoulos $D$-term and a $\theta$-angle; these terms are naturally written in terms of the complex combination $\tau=ir+\frac 1{2\pi}\theta$. The resulting Lagrangian density is thus \begin{eqnarray} {\cal L}&=&\int d^4\theta\left(\|e^{R({V})}\vec{\Phi}\|^2 -\frac1{4e^2}\|{\Sigma}\|^2\right)\nonumber\\ &&+ \left(\int d\theta^+d\theta^- W(\Phi_1,\dots,\Phi_n) + \mbox{ c.c. } \right)\label{GLSM}\\ &&+ \left( {i\over\sqrt 2}\, \int d\theta^+d\bar\theta^- \langle\tau,\Sigma\rangle + \mbox{ c.c. } \right) ,\nonumber \end{eqnarray} where $R=-i\,d\rho:{\bf g}\to{\bf} \def\bf{\bf R}^n$ is the derivative of the representation $\rho$ (with a factor of $-i$ to make it real-valued). Concretely, the general compact abelian group takes the form $G = U(1)^{n-d}\times\Gamma$ where $\Gamma$ is a product of finite cyclic groups. Choosing a basis for the continuous part we have $n{-}d$ vector multiplets $V_a$; the action on the fields is given by $R(V_a)\Phi_i = Q_i^a\Phi_i$ for some integer charges $Q_i^a$. Notice that the discrete group $\Gamma$ does not appear explicitly in the Lagrangian, but it does affect the construction of the field theory---the fields are sections of bundles with structure group $G$. In the case that $\Gamma$ is nontrivial, we recover in this way an orbifold of another theory (cf.\ Ref.~\citelow{phases}). We will be interested in families of such models, parameterized by the coefficients of $W$ and by the instanton factors $q_a := e^{2\pi i\tau_a}$. (There will be a set of complex codimension one in the parameter space along which the model is singular; we will study values of the parameters away from this locus.) A family is thus characterized by the group $G$ and the collection of monomials appearing in $W$. In order to specify these data, it is convenient to introduce a $u\times n$ matrix $P$ of rank $d$ with nonnegative integer entries, and a factorization $P=ST$ of $P$ as a product of integer matrices $S$ and $T$, each of rank $d$. The rows $t_\alpha$ of $T$ can then be used to construct a collection of Laurent monomials $x^{t_\alpha}:=\prod x_i^{t_{\alpha i}}$, and the group $G$ is defined to be the largest subgroup of $H=U(1)^n$ which leaves the monomials $x^{t_\alpha}$ invariant. The monomials $x^{p_r}$ defined by the rows of $P$ are then $G$-invariant by construction, thanks to the relation $p_{ri}=\sum s_{r\alpha}t_{\alpha i}$. Since $p_{ri}\ge0$ by assumption, we may use these monomials to specify the family of interaction polynomials \begin{equation}\label{W} W(x_1,\dots,x_n):={{1\over 2\pi\sqrt 2}}\sum_{r=1}^u c_r x^{p_r}={{1\over 2\pi\sqrt 2}}\sum_{r=1}^u c_r \prod_{i=1}^n x_i^{p_{ri}}\ . \end{equation} Alternatively, if we are given $G$ and a family of polynomials $W$, it is not difficult to reconstruct the matrices $P$, $S$, and $T$. (Actually, $S$ and $T$ are only well-defined up to $(S,T)\mapsto (SL,L^{-1}T)$, with $L$ an invertible integer matrix.) If we start with only $W$, as specified by the rows of the matrix $P$, then the choice of a factorization $P=ST$ amounts to a choice of a subgroup $G\subset U(1)^n$ (which can be less than maximal) under which $W$ is invariant. On the other hand, if we start with only $G$ then the choice of factorized matrix $P=ST$ allows us to specify which of the $G$-invariant monomials should be included in $W$. In particular, it is possible to omit some monomials and in this way to study subsets of the maximal set of all $G$-invariant superpotentials. As we shall see, this possibility is useful. (The assumption made above on the rank of $S$ and $T$ restricts this choice---we must use ``enough'' monomials to get the rank to be correct---but the restriction is not essential and relaxing it would simply require a more cumbersome notation.) Under these conditions, the possible choices for factorizations (and hence for $G$) for a given $P$ are rather limited. In any factorization $P=ST$, the rows of $T$ will generate an integral lattice of rank $d$, and the rows of $P$ will generate a sublattice, also of rank $d$. This implies that the quotient \[\mathop{\rm rowlattice}\nolimits(T)/\mathop{\rm rowlattice}\nolimits(P)\] is a torsion subgroup of ${\bf} \def\bf{\bf Z}^n/\mathop{\rm rowlattice}\nolimits(P)$. There are only finitely many of choices of such a torsion subgroup, once $P$ has been fixed. \section{The $R$-Symmetry} The gauged linear sigma model is not conformally invariant, and will flow to strong-coupling in the infrared. Generically such a theory would be expected to develop a mass gap, but evidence has been found\cite{phases,SilWit} that the theory at the IR fixed point will be nontrivial if we require that the high-energy theory admit a non-anomalous chiral $R$-symmetry. We consider a right-moving $R$-symmetry, acting in such a way that $\theta^+$ has charge $1$ and $\theta^-$ is neutral. Invariance of the kinetic terms requires that the gauge fields ${V}$ be neutral, and their field strengths ${\Sigma}$ hence have charge $1$. Invariance of the superpotential interaction requires that the superpotential have charge $1$. If we let $\mu_i$ denote the $R$-charge of $\Phi_i$ (which may be a rational number) this tells us that \[\sum_{i=1}^n p_{ri} \mu_i=1 \quad \mbox{for all $r$}\ .\] This chiral symmetry can be anomalous in the presence of the gauge fields. A quick computation\cite{phases,summing} shows that the anomaly is given by a function on the Lie algebra proportional to $V\mapsto\mathop{\rm trace}\nolimits(R(V))$; we require that this vanish identically, i.e., that the symmetry be nonanomalous. Since the action of the continuous part of $G$ on the monomial $x_1\cdots x_n$ is via $\exp(\mathop{\rm trace}\nolimits(R(V)))$, this is the same as requiring that $x_1\cdots x_n$ be invariant under the continuous part of $G$ (or in our explicit coordinates that $\sum_i Q_i^a=0$ for all $a$). This in turn will hold exactly when the vector of exponents $(1,\dots,1)$ is a linear combination of the rows of the matrix $T$, with rational coefficients. (If we had wanted $x_1\cdots x_n$ to be invariant under all of $G$, we would have insisted upon integer coefficients.) That is, there is a rational vector $\vec{\lambda}$ such that $\vec{\lambda}^{\rm T} T=(1,\dots,1)$. Since we are assuming that the $d\times u$ matrix $S^{\rm T}$ has rank $d$, it is possible to solve the equation $S^{\rm T}\vec{\nu}=\vec{\lambda}$ for a rational vector $\vec{\nu}$. If we also set $\vec{\mu}=(\mu_1,\dots,\mu_n)^{\rm T}$, then we can write the conditions for an unbroken $R$-symmetry as the existence of $\vec\mu,\,\vec\nu$ such that \begin{eqnarray} P\vec{\mu}&=&(1,\dots,1)^{\rm T}\label{eq1}\\ \vec{\nu}^{\rm T} P&=&(1,\dots,1)\ .\label{eq2} \end{eqnarray} Finally, using the $R$-symmetry and calculating as in Ref.~\citelow{SilWit}, one finds that the central charge $c$ of the fixed-point CFT is determined by \[d-(c/3)=2\sum_{i=1}^n\mu_i=2\,\vec{\nu}\,{}^{\rm T} P\vec{\mu}=2\sum_{r=1}^u\nu_r\ . \] It may be useful to have some examples. \vspace*{0.6cm}\noindent{\normalsize\it Example 1.}\par\vspace*{0.4cm} Consider $1\times1$ matrices $(p)=(s)(t)$ with $s$ and $t$ positive integers. Note that \eqsref{eq1} and (\ref{eq2}) are trivially satisfied in this case. Since $G$ is the group which leaves $x^t$ invariant, we must have $G={\bf} \def\bf{\bf Z}/(t)$. The interaction polynomial is $W(x)=x^p$, so our model is an orbifold of the Landau-Ginsburg version of a minimal model. \vspace*{0.6cm}\noindent{\normalsize\it Example 2.}\par\vspace*{0.4cm} Consider a $u\times 6$ matrix $P$ of rank $5$ such that $p_{0i}=p_{r0}=1$, (letting the $i$ and $r$ indices start at $0$ for this example) implementing \eqsref{eq1} and (\ref{eq2}) directly. This leads to a polynomial of the form \[W(x_0,\dots,x_5)=x_0\,W'(x_1,\dots,x_5)= {{1\over 2\pi\sqrt 2}}\left(c_0\,x_0x_1x_2x_3x_4x_5+x_0\,f(x_1,\dots,x_5)\right),\] where $f$ is a polynomial in $5$ variables containing precisely $u{-}1$ monomials. The existence of a kernel for $p$, an integral generator of which we denote by $(-k,\ell_1,\dots,\ell_5)^{\rm T}$, means that $W$ is invariant under a $U(1)$-action on $(x_0,\dots,x_5)$ with weights $(-k,\ell_1,\dots,\ell_5)$---this implies that $f$ is quasi-homogeneous of degree $k$ and that $k=\sum \ell_i$. One possible factorization $P=ST$ would then be given by taking the rows of $T$ to be a basis for the $U(1)$-invariant Laurent monomials, and expressing the monomials appearing in $f$ in terms of those. By construction, this leads to $G=U(1)$. In general (if $f$ is a generic quasi-homogeneous polynomial) this is the only possible factorization. For special subfamilies (determined by a subset of the rows of the maximal $P$) the polynomial is invariant under additional discrete symmetries (the restriction on $\mathop{\rm rank}\nolimits(P)$ constrains the continuous part), and other factorizations are possible, leading to groups $G$ with a nontrivial discrete part. As we will see below, the weights we have given specify a weighted projective space ${\bf} \def\bf{\bf P}^{(\ell_1,\dots,\ell_5)}$, and $\{f{=}0\}\subset {\bf} \def\bf{\bf P}^{(\ell_1,\dots,\ell_5)}$ defines a (singular) Calabi--Yau hypersurface in that space, closely related to the GLSM built from the data $P=ST$. A particularly interesting case is when $u=6$, and $P$ is $6\times6$. The description of these models in terms of the matrix $P$ was first given by Candelas, de la Ossa, and Katz\cite{cdk} in the course of generalizing the Berglund--H\"ubsch\cite{BHub} construction. \vspace*{0.6cm}\noindent{\normalsize\it Example 3.}\par\vspace*{0.4cm} To make the previous example a bit more concrete, consider a Fermat-type polynomial \[f(x_1,\dots,x_5)=x_1^{a_1}+x_2^{a_2}+x_3^{a_3}+x_4^{a_4} +x_5^{a_5}\] with $\ell_j=k/a_j$ and suppose that $\ell_5=1$. Choosing $G=U(1)$ as in example 2 corresponds to the factorization \[ \left(\begin{array}{cccccc} 1&1&1&1&1&1\\ 1&a_1&0&0&0&0\\ 1&0&a_2&0&0&0\\ 1&0&0&a_3&0&0\\ 1&0&0&0&a_4&0\\ 1&0&0&0&0&a_5 \end{array}\right)= \left(\begin{array}{ccccc} 1&1&1&1&1\\ 1&a_1&0&0&0\\ 1&0&a_2&0&0\\ 1&0&0&a_3&0\\ 1&0&0&0&a_4\\ 1&0&0&0&0 \end{array}\right) \left(\begin{array}{cccccc} 1&0&0&0&0&k\\ 0&1&0&0&0&-\ell_1\\ 0&0&1&0&0&-\ell_2\\ 0&0&0&1&0&-\ell_3\\ 0&0&0&0&1&-\ell_4 \end{array}\right). \] On the other hand, to maximize the group $G$ we should use the entire row space of $P$ as the row space of $T$ which leads to the factorization \[ \left(\begin{array}{cccccc} 1&1&1&1&1&1\\ 1&a_1&0&0&0&0\\ 1&0&a_2&0&0&0\\ 1&0&0&a_3&0&0\\ 1&0&0&0&a_4&0\\ 1&0&0&0&0&a_5 \end{array}\right)= \left(\begin{array}{ccccc} 1&0&0&0&0\\ 0&1&0&0&0\\ 0&0&1&0&0\\ 0&0&0&1&0\\ 0&0&0&0&1\\ k&-\ell_1&-\ell_2&-\ell_3&-\ell_4 \end{array}\right) \left(\begin{array}{cccccc} 1&1&1&1&1&1\\ 1&a_1&0&0&0&0\\ 1&0&a_2&0&0&0\\ 1&0&0&a_3&0&0\\ 1&0&0&0&a_4&0 \end{array}\right). \] This corresponds to an orbifold of the previous model; in fact, this is the quotient found by Greene and Plesser\cite{gp} to correspond to the mirror manifold. Thus we can reproduce the construction of Ref.~\citelow{gp} in this context by appropriate choices of factorization. \section{The Low-Energy Limit} As explained in detail in Ref.~\citelow{phases}, the low-energy limit of this theory can be described explicitly and---when the central charge is an integer---often coincides with a sigma-model on a Calabi--Yau space (for appropriate values of the parameters). To study the low-energy limit one begins by mapping out the space of classical vacua of the theory. To this end, we first solve the algebraic equations of motion for the auxiliary fields $D_a$ (in the vector multiplets) and $F_i$ (in the chiral multiplets) \begin{eqnarray} D_a &=& -e^2\left(\sum_{i=1}^n Q_i^a |\phi_i|^2 - r_a\right)\label{D}\\ F_i &=& -{\partial W\over\partial\phi_i}\ .\label{F} \end{eqnarray} The potential energy for the bosonic zero modes is then \[U = {1\over 2e^2}\sum_{a=1}^{n-d}D_a^2 + \sum_{i=1}^n |F_i|^2 + \sum_{a,b}\bar\sigma_a\sigma_b\sum_{i=1}^n Q_i^aQ_i^b|\phi_i|^2\ ,\] where $\phi_i,\,\sigma_a$ are the lowest components of $\Phi_i,\,\Sigma_a$ respectively. The space of classical vacua is the quotient by $G$ of the set of zeros of $U$. Neglecting for a moment the superpotential, the space of solutions (setting $\sigma=0$) is $D^{-1}(0)/G$. For values of the instanton factors $q_a$ in a suitable range this is a toric variety of dimension $d$. In the ``geometric'' case in which $\vec{\mu}$ has $N$ $1$'s and $n{-}N$ $0$'s, this variety can be recognized as the total space of a sum of $N$ line bundles over a {\it compact}\/ toric variety of dimension $d-N$. The equations $F_i=0$ then determine (for generic values of the parameters $c_r$) a complete intersection subvariety of codimension $N$ in the base space, homologous to the intersection of the divisors associated to the $N$ line bundles. The condition in \eqref{eq2} implies that this subvariety is Calabi--Yau. The fixed-point CFT is given by the nonlinear sigma model with this target space. The moduli of this CFT are the marginal operators in \eqref{GLSM}. Na\"{\i}vely, both the $q_a$ and the $c_r$ would appear to be marginal couplings. The latter, however, are not all independent. As is well known, some of them can be absorbed in rescalings of the fields $\Phi_i$. The true moduli are thus the $q_a$ and the scaling-invariant combinations of the $c_r$. There will in general be other marginal operators in the model which do not appear explicitly in the Lagrangian; we restrict our attention to the subspace of those that do. Of course, the linear (or more accurately, toric) structure with which we have endowed our moduli space is an artifact of our description. In particular, it is consistent with the natural complex structure that this moduli space carries intrinsically, but bears no relation to the K\"ahler structure determined by the Zamolodchikov metric. In other words, the $q_a$ and $c_r$ are not the ``special'' coordinates on this space. For example, in example 2 above in which $G=U(1)$, the $D$-term equation is \[-k\,|x_0|^2+\sum_{i>0}\ell_i|x_i|^2 - r=0,\] so when $r>0$ we cannot have $x_i=0$ for all $i>0$. The space of solutions $D^{-1}(0)/G$ can be recognized as the total space of the canonical bundle over ${\bf} \def\bf{\bf P}^{(\ell_1,\dots,\ell_5)}$. If we impose the $F$-term equation as well \[\sum_{i=0}^5\left|\frac{\partial W}{\partial x_i}\right|^2 =|W'|^2+|x_0|^2\sum_{i=1}^5\left|\frac{\partial W'}{\partial x_i}\right|^2 =0,\] then for a generic choice of the coefficients of $W$ (away from a codimension-one subspace we avoid as promised above) the space of solutions is given by $x_0=0$ and $W'(x)=0$, yielding a Calabi--Yau hypersurface in ${\bf} \def\bf{\bf P}^{(\ell_1,\dots,\ell_5)}$. Further study shows that the low-energy excitations are tangent to this, so that the low-energy theory is a nonlinear sigma model with this Calabi--Yau target space. \section{Mirror Symmetry} We are now in a position to state the mirror symmetry conjectures for this class of models. Mirror symmetry relates two Calabi--Yau manifolds which lead to isomorphic CFT's when used as target spaces for supersymmetric nonlinear sigma models. The mirror isomorphism reverses the sign of the right-moving $R$-symmetry; many of the fascinating properties of mirror pairs can be traced to this feature. Given a GLSM family determined by $P=ST$ we will construct the mirror family in the same class of models. We incorporate the sign change on $R$ by writing the dual model using {\it twisted}\/ chiral matter fields coupled to twisted gauge multiplets (with {\it chiral}\/ field strengths). We will call such a model a {\it twisted}\/ gauged linear sigma model. As above, we characterize the family of models by a factorized matrix of nonnegative integers $\widehat P = \widehat S\widehat T$. We use this data to construct a twisted superpotential $\widehat W$ and write a Lagrangian density (compare \eqref{GLSM}) \begin{eqnarray} {\cal L}&=&\int d^4\theta\left(-\|e^{\widehat R({\widehat V})} \widehat {\vec{\Phi}}\|^2 +\frac1{4e^2}\|{\widehat \Sigma}\|^2\right)\nonumber\\ &&+ \left(\int d\theta^+d\bar\theta^- \widehat W({\widehat\Phi}) + \mbox{ c.c. } \right)\label{TGLSM}\\ &&+ \left( {i\over\sqrt 2}\, \int d\theta^+d\theta^- \langle\widehat \tau,\widehat \Sigma\rangle + \mbox{ c.c. } \right) ,\nonumber \end{eqnarray} where $\widehat\Sigma = -{1\over 2\sqrt 2}\bar D_+\bar D_-\widehat V$ is the (chiral) field strength. The {\it mirror conjecture}\/ states that if we set \begin{eqnarray} \widehat P &=& P^{\rm T}\nonumber\\ \widehat S &=& T^{\rm T}\label{mirp}\\ \widehat T &=& S^{\rm T}\nonumber \end{eqnarray} the two families of CFT's defined by the infrared limits of the two linear models are isomorphic. This is essentially just a translation of the conjecture of Batyrev and Borisov\cite{batbor}, versions of which have appeared in Refs.~\citelow{ag:rigid,cdk}. In fact, the statement given here generalizes the conjecture somewhat since the factorization makes possible a map between subfamilies. One can make a more precise statement of the conjecture---specifying explicitly the map between the two parameter spaces which relates isomorphic low-energy limits \begin{eqnarray} q_a &=& \prod_{r=1}^u \widehat c_i^{Q_i^a}\nonumber\\ \widehat q_{\widehat a} &=& \prod_{i=1}^n c_r^{\widehat Q_r^{\widehat a}}\ .\label{mondiv} \end{eqnarray} Note that as expected only the true moduli appear on the right-hand side of \eqref{mondiv}. The asymptotic limits of \eqref{mondiv} constitute the {\it monomial-divisor mirror map}, first proposed in Refs.\ \citelow{catp,mondiv} (see also Ref.~\citelow{Bat:qu}) and extended in Refs.\ \citelow{ag:rigid,multdiv}. The ability to write a global version of this map and not just an asymptotic form is related to the coordinates we use on the moduli space (for a full discussion of this including a proof of \eqref{mondiv} for a class of examples see Ref.~\citelow{summing}). We note that this statement of the conjecture is very reminiscent of recent results on duality in four-dimensional supersymmetric gauge theories\cite{SWI,SWII,S}. Two distinct gauge theories with nontrivial dynamics lead to the same low-energy physics; further, instanton effects in one model are reproduced by classical physics in the other. In example 3 above, it should be clear that this construction interchanges the two choices of factorization, leaving the (symmetric) matrix $P$ unchanged. As mentioned above, this reproduces the orbifold construction of the mirror in Ref.~\citelow{gp}, clearly and explicitly demonstrating how the general construction reduces to this upon restricting to a subfamily of all possible $W$---precisely the subfamily invariant under the maximal discrete group. \section{Abelian Duality} The formulation we have given for the mirror conjecture in the context of the GLSM immediately suggests a relation to abelian duality. We recall that given a theory with an abelian continuous global symmetry, one can use duality to obtain an equivalent theory. The dual model is obtained by gauging the global symmetry, introducing Lagrange multipliers which constrain the connection to be pure gauge (ensuring that the theory is actually unchanged). One must then eliminate the original degrees of freedom, in the process generating a nontrivial effective action for the Lagrange multipliers, which become the fundamental degrees of freedom for the dual model. Classically this is accomplished by gauge-fixing. In a theory with $N{=}2$ supersymmetry and chiral charged fields, the Lagrange multipliers mentioned above will be {\it twisted\/} chiral fields, and will appear in the twisted superpotential as $\widehat\Lambda\Sigma$. Thus the dual variables will naturally be twisted chiral\cite{GHR}, as expected for the mirror model. This idea was pursued in Refs.\ \citelow{G,RV,GW,BH}, and in special cases leads to an interpretation of mirror symmetry as abelian duality. In the general case, however, the approach runs into difficulties. The most obvious one is that a generic Calabi--Yau manifold $M$ has no isometries, hence the associated two dimensional CFT lacks suitable global symmetries. If this approach is to lead to the equivalence of GLSM models which is tantamount to the mirror conjecture, some modification will be required. The two linear models are {\it not\/} equivalent; the conjecture implies only that they become equivalent in the extreme low-energy limit. This is consistent with the fact that the model has nontrivial dynamics. As discussed above this is similar to the recent discoveries in four-dimensional supersymmetric gauge theories. In the context of the GLSM, there is a natural symmetry group one would attempt to use. This is the group $H = U(1)^n$ acting by phases on the fields $\Phi_i$. The difficulty, of course, is that this symmetry is explicitly broken by the superpotential \eqref{W} to the subgroup $G$. Since this symmetry is gauged in \eqref{GLSM}, the resulting model has no global symmetries at all (except the $R$-symmetry which we cannot gauge).\footnote{This once led E. Witten to describe the problem of understanding mirror symmetry in this model as the question ``How to perform duality on a non-symmetry?''} We can attempt to overcome this obstacle by recasting the symmetry-breaking terms as anomalies. To implement this idea, introduce a set of twisted chiral fields $\widehat\Psi_r$ into the model, coupled to twisted gauge multiplets $\widehat V_r$ gauging the group $\widehat H = U(1)^u$ which acts by phases, i.e.\ consider the additional term in the kinetic energy \begin{equation} {\cal L}_{\widehat k} = \int d^4\theta \left( -\|\widehat{\vec\Psi} e^{\widehat{V}}\|^2 +{1\over 4e^2}\|{\widehat{\Sigma}}\|^2\right)\ . \label{khat} \end{equation} We then replace the superpotential \eqref{W} by \begin{equation} W_s = {{1\over 2\pi\sqrt 2}} \sum_{r=1}^u{\widehat{\Sigma}}_r\left[\log\left( c_r\prod_{i=1}^n\Phi_i^{p_{ri}}\right)+1\right]\ . \label{Ws} \end{equation} In this form it is clear that $H$ is broken by $\widehat H$ anomalies, as desired. Note that this interaction does not break the $R$-symmetry. However, the couplings $c_r$ have nonzero beta functions, which will cause them to grow large, so the $\widehat H$ gauge theory is strongly coupled at low energies. In this limit, as discussed in Refs.\ \citelow{phases,summing}, this sector of the model is in a confining phase, in which the lowest component $\widehat\sigma$ of ${\widehat{\Sigma}}$ gets a nonzero expectation value. The charged fields are then all massive, with masses of order this expectation value, and the light degrees of freedom are in the ${\widehat{\Sigma}}$ multiplet. The effective action for these can be reliably computed at one-loop order. Integrating out the $\widehat\Psi$'s leads to the effective superpotential \begin{equation} W_{\rm eff} = {{1\over 2\pi\sqrt 2}}\sum_{r=1}^u {\widehat{\Sigma}}_r\left[\log\left( c_r\prod_{i=1}^n\Phi_i^{p_{ir}}\right)+1 - \log{\widehat{\Sigma}}_r\right]\ . \label{Wsef} \end{equation} In the effective theory we can treat ${\widehat{\Sigma}}$ as a chiral field; since there are no charged fields we can forget its relation to a gauge symmetry. At low energy, we can eliminate ${\widehat{\Sigma}}$ from \eqref{Wsef} to get ${\widehat{\Sigma}}_r = c_r\prod_{i=1}^n\Phi_i^{p_{ir}}$. When substituted in \eqref{Wsef}, this leads to \eqref{W} as the effective superpotential for $\Phi$. Thus at low energies this model is equivalent to the original GLSM, while the symmetry-breaking terms are explicitly exhibited as anomalies. We now wish to perform a duality transformation with respect to the anomalous symmetry. Gauging an anomalous symmetry appears inconsistent, but in performing a duality transformation an anomalous abelian symmetry can be restored by assigning transformation properties (under the twisted symmetries) to the Lagrange multipliers.\cite{GR} In the case at hand, gauging the global symmetry means that the gauge group becomes all of $H$. We should also introduce Lagrange multipliers, transforming under $\widehat H$, to constrain the field strength to lie in $\bf g\subset\bf h$. Classically, integrating these out will reproduce the original model. This is not true quantum mechanically, however, as is most easily seen by introducing a small kinetic term for the Lagrange multipliers, and considering the limit as this term is removed. It then becomes clear that the contribution of these to the one-loop effective superpotential for ${\widehat{\Sigma}}$ is constant in the limit. Introducing these extra fields would thus destroy the one-loop calculation that led to \eqref{Wsef}. This suggests that we consider a related Lagrangian, in which the twisted chiral matter is an accurate ``reflection'' of the chiral matter. This defines a model that is manifestly mirror-symmetric. Our proposal for this new model contains chiral fields $\vec\Phi$ and twisted chiral fields $\widehat{\vec\Phi}$, and (twisted) gauge multiplets gauging the entire group $H$ ($\widehat H$). We begin with the Lagrangian density \begin{eqnarray} {\cal L} &=& \int d^4\theta \left( \sum_i \left( J_i - {1\over 4e^2}|\Sigma_i|^2\right) - \sum_r\left( \widehat J_r - {1\over 4e^2}|{\widehat{\Sigma}}_r|^2\right) + {1\over 2\pi}\sum_{ri}p_{ri} \log J_i \log \widehat J_r \right)\nonumber\\ && + \left({{i\over \sqrt 2}}\int d\theta^+d\theta^-\langle{\widehat{\Sigma}},\widehat\tau\rangle +{\rm c.c}\right) + \left({{i\over \sqrt 2}}\int d\theta^+d{\bar\theta}^-\langle\Sigma,\tau\rangle +{\rm c.c}\right) \ ,\label{our} \end{eqnarray} where $J_i = |\Phi_ie^{V_i}|^2$ and likewise for $\widehat J_r$. This action is manifestly invariant under $H\times\widehat H$. It is classically equivalent to the action we obtain by integration by parts, \begin{eqnarray} {\cal L_{\rm eq}} &=& \int d^4\theta \left( \sum_i \left( J_i - {1\over 4e^2}|\Sigma_i|^2\right) - \sum_r\left( \widehat J_r - {1\over 4e^2}|{\widehat{\Sigma}}_r|^2\right) + {2\over \pi}\sum_{ri}p_{ri} V_i\widehat V_r\right)\nonumber\\ && + \left(\int d\theta^+d\theta^- W({\widehat{\Sigma}},\Phi)+{\rm c.c}\right) + \left(\int d\theta^+d{\bar\theta}^- \widetilde W(\Sigma,{\widehat\Phi}) +{\rm c.c}\right) \ ,\label{ourpi} \end{eqnarray} where the modified superpotentials are \begin{eqnarray} W({\widehat{\Sigma}},\Phi) &=& {{1\over 2\pi\sqrt 2}}\sum_{r=1}^u {\widehat{\Sigma}}_r\left[\log\left( \widehat q_r\prod_{i=1}^n\Phi_i^{p_{ri}}\right)+1\right] \nonumber\\ \widetilde W(\Sigma,{\widehat\Phi}) &=& {{1\over 2\pi\sqrt 2}}\sum_{i=1}^n \Sigma_i\left[\log\left( q_i\prod_{r=1}^u{\widehat\Phi}_r^{-p_{ri}}\right)+1\right]\ ,\label{Wt} \end{eqnarray} We see that the conditions for a nonanomalous $R$-symmetry are precisely \eqsref{eq1} and (\ref{eq2}). When these hold we expect to find at low energies an $N{=}2$ superconformal field theory where the $R$-symmetry defined above becomes the chiral $U(1)$ contained in the superconformal algebra. We will use the second formulation as our definition of the quantum theory. (In the presence of instantons the integration by parts requires the consideration of boundary terms, which are nontrivial.) It is worthwhile noting the way in which \eqref{ourpi} manages to be gauge-invariant despite the manifestly non-invariant interactions. Under a gauge transformation (in $H\times\widehat H$) the variation of \eqref{Wt} is cancelled, up to a total derivative, by the variation of the $V\widehat V$ term. Thus the full action is invariant under gauge transformations approaching the identity at infinity, while transformations with constant parameter are still anomalous. We note that this cancellation holds precisely when the exponents in the two superpotentials are related as in \eqref{Wt}. In a derivation of mirror symmetry along these lines this would be the origin of the first part of \eqref{mirp}. At this point one still needs to show that the theory defined by \eqref{ourpi} is equivalent to \eqref{GLSM}. This can be achieved using the fact that the coefficients $q_i,\,\widehat q_r$ once more contain redundant deformations which can be undone by field rescalings (in fact this is precisely the holomorphic extension of the statement above about anomalies). One can perform a redundant deformation of the model to find a region in parameter space in which the ${\widehat{\Sigma}},{\widehat\Phi}$ system is in a confining phase and the one-loop approximation that led to \eqref{Wsef} is valid. One can also show that in this region the integration over ${\widehat\Phi}$ will lead to constraints on $\Sigma$, restricting the gauge group to $G$. In fact, the last two equations of \eqref{mirp} arise naturally in this context as well. The manifest mirror symmetry would then lead one to expect that \eqref{ourpi} will also be equivalent to \eqref{TGLSM} under the conditions \eqsref{mirp} and (\ref{mondiv}). To see this one performs redundant deformations to a region in parameter space in which the $\Sigma,\Phi$ system confines and follows a ``mirror'' version of the argument sketched above. This naturally incorporates all of the conditions in \eqsref{mirp} and (\ref{mondiv}). We hasten to point out however that \eqref{ourpi} is not in fact mirror symmetric. The reason for this is the sign in the second of \eqref{Wt}.\footnote{This sign was inadvertently omitted in earlier stages of this work, leading to conclusions---reported in several talks by the authors---differing from those presented here.} Unexpectedly, this sign, which is required for gauge invariance since it follows by integration by parts from \eqref{our}, cannot be removed (it is related to the relative sign of the kinetic terms for the chiral and twisted chiral fields). Similar signs appear often in discussions of duality, and are usually harmless. In the case at hand, however, the sign multiplies a logarithm, and its effect is to lead, at least superficially, to a twisted superpotential polynomial in ${\widehat\Phi}^{-1}$ rather than ${\widehat\Phi}$ after performing the ``mirror'' argument. This would seem to shatter the hopes for an understanding of mirror symmetry for these models along the lines presented here. However, the model we present does seem to incorporate naturally many of the features required for a derivation of the conjectures in section four. It is possible that some modification of this model will lead to the desired result. Future study will tell if this is indeed the case. \vspace*{0.6cm} \noindent{\normalsize\bf Acknowledgements} \vspace*{0.4cm} We thank P. Candelas and X. de la Ossa for many discussions and for collaboration in the early stages of this work, and P. Argyres, P. Aspinwall, B. Greene, G. Moore, M. Ro\v cek, and especially E. Witten and N. Seiberg for discussions. The work of D.R.M.\ was supported in part by the United States Army Research Office through the Mathematical Sciences Institute of Cornell University, contract DAAL03-91-C-0027, and in part by the National Science Foundation through grants DMS-9401447 and PHY-9258582. The work of M.R.P.\ was supported by National Science Foundation grant PHY-9245317 and by the W.M. Keck Foundation. \vspace*{0.6cm} \noindent{\normalsize\bf References}
\section{INTRODUCTION} Integrable models in two space-time dimensions provide a fascinating arena for investigating non-perturbative phenomena in quantum field theory \cite{Raj,Col,AAR}. The integrable theories that we shall focus on here can be divided into three broad classes, but they all have in common the properties of asymptotic freedom and dynamical mass generation which make them akin to realistic models of particle interactions in four dimensions. First, there are bosonic sigma models based on a symmetric space $G/H$. Any model of this sort is classically integrable, but unfortunately anomalies can destroy integrability at the quantum level unless $H$ is simple. Second, there are fermionic models of Gross-Neveu or Thirring type with four-fermion interactions. Finally, there are hybrids of the two previous classes in which fermions are added to bosonic symmetric space models in various ways. In some cases the addition of fermions can cancel the anomalies encountered in the purely bosonic theory, leading to interesting new examples of quantum integrable theories; in particular this is believed to occur for supersymmetric sigma models based on a symmetric space $G/H$. A review of these matters with detailed references can be found in \cite{AAR}. Exact S-matrices have been proposed for many of the two-dimensional models which are thought to be integrable at the quantum level \cite{AAR,ZZ,SMGN,SMCGN,SMPCM,SW,SMCPN}. These S-matrices describe the scattering of some conjectured set of particles states, and they are postulated on the basis of the specific symmetries of the model in question together with the usual axioms of S-matrix theory and the powerful constraint of factorization. (There may also be additional information available about the theory, such as the existence of bound-states.) However, such S-matrices are always subject to CDD ambiguities \cite{CDD,ZZ} which cannot be further constrained by these general considerations. What is needed is some completely non-perturbative way of fixing the CDD ambiguities by testing the proposed equivalence between an S-matrix on the one hand and renormalized Lagrangian perturbation theory on the other. In this paper we shall explain such a programme which can be applied to any of the types of theory mentioned above. The technique, which was pioneered in \cite{HMN,HN,FNW} following earlier work in \cite{TBA,PW,W,JNW,ZTBA}, can be applied to any integrable model possessing a group $G$ of global symmetries. The particle states in such a theory fall into representations of $G$, and the symmetry is generated by conserved charges in the Lie algebra. The idea is to couple the theory to some particular conserved charge $Q$ by modifying the Hamiltonian from $H$ to $H-hQ$, where $h$ is a coupling constant of mass dimension 1 (so that $Q$ is dimensionless). The corresponding change in the ground-state energy density $\delta {\cal E}(h)={\cal E}(h)-{\cal E}(0)$ can be computed straightforwardly in perturbation theory. But in the special case of an integrable model, it can also be computed by a very different non-perturbative method starting from the exact S-matrix and using the Thermodynamic Bethe Ansatz (TBA). The comparison of these calculations provides a powerful check that the S-matrix and the Lagrangian really do correspond to one another and it also allows one to extract an {\it exact\/} expression for the mass gap of the model. We now describe more quantitatively how this comparison works. The result of a perturbative calculation of the ground-state energy density---which is of course a renormalization group-invariant quantity---is an expansion \[{ \delta {\cal E}(h) = h^2 \sum_{j=0}^\infty {\alpha}_{j} \, g(h/\Lambda)^{j-1} }\] in ascending powers of the running coupling $g(h / \Lambda)$ where the $\alpha_j$ are dimensionless numbers. The running coupling can be found by integrating the usual beta-function equation \[ \mu \, {d g \over d \mu} = \beta(g)= - \sum_{j=1}^\infty \beta_j \, g^{j+1} \quad {\hbox{\rm to~obtain}} \quad {1 \over g(\mu / \Lambda)} = \beta_1 \ln {\mu \over \Lambda} + {\beta_2 \over \beta_1} \ln \ln {\mu \over \Lambda} + {\cal O} {\hbox{\Large (}} \ln \ln {\mu \over \Lambda} {\hbox{\Large /}} \ln {\mu \over \Lambda} {\hbox{\Large )}} \, , \] where the absence of a constant term in the solution defines the mass scale $\Lambda$ for the particular renormalization scheme used. In the models we are considering, $\beta_1 > 0$ and we expect perturbation theory to be valid in the asymptotic regime where $\mu\gg\Lambda$. Combining the expressions above we find that the ground-state energy density is given by: \begin{equation} {\delta {\cal E}(h) \over h^2} = \alpha_{0}\beta_1\ln{h\over\Lambda} +\alpha_{0}{\beta_2\over\beta_1} \ln\ln{h\over\Lambda}+\alpha_1 + {\cal O} {\hbox{\Large (}} \ln \ln {h \over \Lambda} {\hbox{\Large /}} \ln {h \over \Lambda} {\hbox{\Large )}} \, .\label{eq;E1} \end{equation} Notice that if $\alpha_0 \neq 0$, there is a {\it classical\/} or {\it tree-level\/} contribution to the ground-state energy density and this quantity is consequently unbounded as $h$ becomes large. Otherwise, $\alpha_0 = 0$ and the leading contribution to the ground-state energy is then the constant term $\alpha_1$. We shall have more to say shortly about the important differences between these situations. Turning now to the TBA calculation: in order to extract the ground-state energy from the S-matrix, one must confront a set of coupled integral equations of Wiener-Hopf type. The special circumstances of interest to us are those in which the temperature is zero, with the coupling to the charge $Q$ acting like a chemical potential. In general the resulting equations cannot be solved exactly, but it is possible, at least under certain simplifying assumptions, to generate an expansion of the solution in $h/m$ valid when $h{\gg}m$, where $m$ is some physical mass parameter occurring in the S-matrix. The result is \begin{equation} {\delta {\cal E}(h) \over h^2} = \kappa_0\ln{h\over m}+\kappa_1\ln\ln{h\over m}+\kappa_2 + {\cal O} {\hbox{\Large (}} \ln \ln {h \over m} {\hbox{\Large /}} \ln {h \over m} {\hbox{\Large )}} \, ,\label{eq;E2} \end{equation} where the $\kappa_j$ are dimensionless numbers. The exact values of the parameters $\kappa_j$ depend sensitively on the analytic structure of the S-matrix and, in particular, one finds that the presence or absence of CDD factors alters dramatically the result for the ground-state energy obtained in this way. Equality of the expressions (1) and (2) gives a powerful check that the Lagrangian and S-matrix descriptions are consistent. The greater the accuracy to which we can calculate these expressions, the more conclusive this check will be. Even finding just the first few terms, however, can provide strong evidence in favour of the proposed S-matrix used in the TBA calculation, allowing us to argue that any alteration by CDD factors would destroy the delicate agreement with the perturbative result. In addition, the consistency of (1) and (2) clearly determines the mass gap $m / \Lambda$, at least if we have calculated the expressions involved to sufficient accuracy. At this stage it is useful to look more closely at the two possible cases to which we drew attention earlier. If $\alpha_0 \neq 0$, we see that the TBA calculation must reproduce the first two coefficients of the beta-function through the conditions: \begin{equation} \alpha_0 \beta_1 = \kappa_0 \ , \qquad \alpha_0 \beta_2 / \beta_1 = \kappa_1 \ , \end{equation} providing a highly non-trivial test of the S-matrix. In these circumstances we can also read off the value of the mass gap in the model: \begin{equation} \ln (m/\Lambda) = (\kappa_2 - \alpha_1)/\kappa_0 = (\kappa_2 - \alpha_1) / \alpha_0 \beta_1 \, . \end{equation} We emphasize that these relations are deduced by comparing the terms written {\it explicitly\/} in (1) and (2) and they therefore rely on just a one-loop perturbative calculation of the ground-state energy. If, instead, $\alpha_0 = 0$, we must have $\kappa_0 = \kappa_1 = 0$ and $\kappa_2 = \alpha_1$ for consistency and we are then unable to find the mass-gap by taking the expressions to the order given explicitly in (1) and (2). To extract $m/\Lambda$ in this situation one needs to extend both the TBA and perturbative calculations to higher orders; in fact it is not hard to see that this would involve at least a three-loop perturbative calculation. The strategy we have just outlined has been applied to several series of integrable models, with results we shall summarize in section 3. The expressions for the exact mass-gaps derived in this way are very useful, because they provide bench-marks for the reliability of other non-perturbative approaches, such as lattice simulations. To minimize the amount of work involved, it is clearly desirable to try to choose $Q$ so as to produce a classical term in the perturbative expression for the ground-state energy (1) because, as we have explained, the mass-gap can then be found from the TBA analysis in conjunction with a one-loop perturbative calculation. It can be shown \cite{EH3} that such a choice of $Q$ is possible for any bosonic or supersymmetric sigma model based on a symmetric space $G/H$, a result which unifies the treatment of various examples considered previously in \cite{HMN,HN,BNNW,THIII,EH1,EH2}. For the purely fermionic Gross-Neveu models, however, it seems that generically the classical contribution vanishes, and so a three-loop calculation is necessary in order to find the mass gap \cite{FNW,CGN}. There is another crucial consideration to be borne in mind when choosing $Q$. In each of the cases considered so far, an important simplifying assumption was made regarding the TBA calculation, namely, that for very particular choices of the charge $Q$, only a small number of particles---in fact those with the largest charge/mass ratio---contribute to the new ground-state. The original one-particle states can be chosen to be eigenvectors of the new Hamiltonian $H-hQ$ with eigenvalues $m_i\cosh\theta_i-hq_i$, where $q_i$ is the charge of the particle labelled by $i^{\rm}$ and $\theta_i$ is its rapidity. If $i=1$ labels the particle type with the largest charge/mass ratio, then for sufficiently small values of $h$ such that $hq_1<m_1$ it is not energetically favourable to find any particles in the new ground-state, and so $\delta {\cal E}(h)=0$. As $h$ passes the threshold value $m_1/q_1$, it suddenly becomes favourable to fill the original vacuum with particles. For large $h$, it would seem that any arbitrary particle could appear in the new ground-state and hence the TBA calculation would have to keep track of all particles. However, for very particular choices of the charge $Q$ it seems that the particles with largest charge/mass ratio actually repel other particles and so only they appear in the ground state. This assumption greatly simplifies the solution of the TBA equations and in most of the existing papers it is taken as a working hypothesis which is vindicated by the consistency of the final results. With more care, it can actually be proven from an analysis of the full TBA equations \cite{EH3}. Having introduced the general idea behind the technique, we shall discuss in the next section how the TBA equations can be derived under the simplifying assumption explained above. We shall then summarize the known results, concluding with a more detailed example which illustrates the most important points. \section{THE TBA EQUATIONS} In the cases where a single particle-type contributes to the ground state, the TBA analysis is rather straightforward. The elastic scattering of two particles of the same species is described by an S-matrix element which is just a phase $S(\theta)$, where $\theta$ is the rapidity difference of the incoming particles. Since particle number and momenta are conserved on interaction, it makes sense to consider single particle states. In the dilute regime, where the particles are on average well separated, the wavefunction of $N$ particles is built up from these single particle states and has the form \[ \Psi(x_1,\ldots,x_N)=\sum_{Q\in S_N}\Theta(x_Q)\zeta(Q) \exp im (x_1\sinh\theta_1 + \ldots + x_N\sinh\theta_N ) \] where the sum is taken over all permutations $Q=\{Q_1,\ldots,Q_N\}$ of $\{1,\ldots,N\}$ with \[ \Theta(x_Q)=\left\{\begin{array}{ll}1&{\rm if}\ x_{Q_1}<x_{Q_2}<\cdots<x_{Q_N}\\ 0&{\rm otherwise}\end{array}\right. \] and where $\zeta(Q)$ are numbers defined via the S-matrix so that if $Q$ differs from $Q'$ only by an exchange of two specific elements of the list $Q_i$ and $Q_j$, say, then \[ \zeta(Q')=S(\theta_i-\theta_j)\zeta(Q) \ . \] This means that the $\zeta(Q)$ are determined up to an overall factor. We impose periodic boundary conditions on the wavefunctions: \[ \Psi(x_1,\ldots,x_j,\ldots,x_N)=\Psi(x_1,\ldots,x_j{+}L,\ldots,x_N)\ \] which imply that, for each fixed $j$, \[ \exp({imL\sinh\theta_j})\prod_{k\neq j}S(\theta_j-\theta_k)=1\ . \] Taking the logarithm of these equations we have \begin{equation} mL\sinh\theta_j-i\sum_{k\neq j}\ln S(\theta_j-\theta_k)=2\pi n_j \end{equation} where the $\{n_j\}$ are a set of $N$ integers which we assume determine uniquely the set of rapidities $\{\theta_j\}$ (this is guaranteed in the dilute regime where the $\sinh\theta_j$ dominates in (5)). It is crucial that in all the integrable models we are considering here the particles behave as fermions in rapidity space. The condition (5) then proves to be the key ingredient when we come to analyse the thermodynamics of such a system. The thermodynamic limit can be taken in the usual way by letting $N\rightarrow\infty$ and $L\rightarrow\infty$ with $N/L$ fixed. It now becomes meaningful to talk about the density of one-particle states in rapidity space, $\varrho (\theta)$, defined by $dn=L\varrho(\theta)d\theta$, and also the density of {\it occupied\/} states in rapidity space, $\sigma(\theta)$. It is also convenient at this stage to define from the S-matrix the quantities $K(\theta)$ and $R(\theta)$ given by \begin{equation} K(\theta) = \delta(\theta) - R(\theta)= {1\over2\pi i}{d\over d\theta}\ln S(\theta) \, . \end{equation} By differentiating the key condition (5) we deduce \begin{equation} \varrho(\theta)={m\over2\pi}\cosh\theta+K*\sigma(\theta) \end{equation} (where $f*g(\theta)=\int_{-\infty}^\infty d \theta' f(\theta') g(\theta-\theta')$ for any two functions $f$ and $g$). When $K(\theta)=0$, (7) reduces to the usual expression for the density-of-states in a free theory. But when there is an interaction between the particles, (7) implies that the total density and the density of occupied states are coupled in a complicated manner. Of course for overall consistency we must have $\sigma(\theta)\leq\varrho(\theta)$, which is guaranteed in the dilute regime since $\sigma(\theta)$ is then small compared with $\varrho(\theta)$. For a macrostate specified by the density $\sigma(\theta)$, the value of $H-hQ$ per unit length is \[ {\cal E}(\sigma;h)= \int_{-\infty}^\infty {d\theta\over2\pi}(m\cosh\theta-h)\sigma(\theta) \, . \] (where we assume the charge is normalized to unity on our preferred particle states). To find the ground-state energy density we must minimize ${\cal E}(\sigma;h)$ with respect to $\sigma(\theta)$ subject to the constraint (7). The solution to this variational problem involves filling all the available states of lowest rapidity first, so $\sigma(\theta)=\varrho(\theta)$ up to some Fermi rapidity $\pm\theta_{\rm F}$. It can be shown that this problem can be solved in terms of an energy density $\epsilon(\theta)$ which satisfies the TBA equation at zero temperature: \begin{equation} \epsilon^+(\theta)+R*\epsilon^-(\theta)=m\cosh\theta-h \end{equation} with the notation \[ f^\pm(\theta)=\left\{\begin{array}{ll} f(\theta)&f(\theta){>\atop<}0\\ 0&{\rm otherwise}.\\\end{array}\right. \] The solution to (8) is concave and negative between the values $\pm\theta_{\rm F}$, which are fixed by the condition $\epsilon(\pm\theta_{\rm F})=0$. The final result for the way in which the ground-state energy density behaves as a function of $h$ is then given simply by \begin{equation} \delta {\cal E}(h)={m\over2\pi}\int_{-\theta_{\rm F}}^{\theta_{\rm F}}d\theta \, \cosh \theta \, \epsilon(\theta) \ . \end{equation} The problem of finding the ground-state energy density has thus been reduced to solving the Wiener-Hopf integral equation (8). It is not possible to find the solution exactly for arbitrary $h$. But we require the solution only in the asymptotic regime with $h\gg m$ and it is explained in \cite {FNW} and \cite{BNNW} how to develop such an asymptotic expansion for the solution based on the original approach of \cite{JNW}. The nature of the expansion depends crucially on whether or not the Fourier transform $\hat R(\omega)$ of the kernel $R(\theta)$ defined via \[ R(\theta)=\int_0^\infty{d\omega\over\pi}\cos(\omega \theta)\hat R(\omega) \, , \] vanishes at the origin. Suppose that $\hat R(0)=0$, and that we can decompose $1/ \hat R (\omega) = G_+(\omega) G_-(\omega)$ where $G_{\pm}(\omega)$ are analytic in the upper/lower half planes with $G_+(\omega) = G_- (-\omega)$. It can be shown that if $G_+(i\xi)$ has an expansion for small $\xi$ like \begin{equation} G_+(i\xi)={k\over\sqrt\xi}\exp(-a\xi\ln\xi)\left(1-b\xi+{\cal O}(\xi^2)\right) \end{equation} then the ground-state energy density for $h\gg m$ takes the form (2) with \begin{equation} \kappa_0 = -k^2/4 \ , \quad \kappa_1/\kappa_0 = a + 1/2 \ , \quad {\kappa_2 / \kappa_0 = \ln (\sqrt{2\pi}/G_+(i)) - 1} +\ln k +a(\gamma_{\rm E}-1+\ln8)-b \ . \end{equation} This matches the expansion in perturbation theory {\it with\/} a classical term in (1). If $\hat R(0)\neq0$, on the other hand, then the asymptotic expansion proceeds slightly differently and the result is an expression of the form (2) with $\kappa_0 = \kappa_1 =0$. This matches the perturbative result (1) in the case when there is no tree-level contribution. At this point we should remark that in the regime $h\gg m$ considered above, the Fermi rapidity is large and so the ground-state will contain a large number of particles. It would seem therefore that the system is far from the dilute situation used in our naive derivation of the TBA equations. It is one of the miracles of the TBA that it seems, nevertheless, to be valid for many systems in the deep ultra-violet regime, although the exact reason for this is not understood. The simple TBA analysis which we have presented also assumed that there is only one particle type that contributes to the new ground-state. In general many particles with different quantum numbers and masses can contribute and the TBA analysis then becomes much more complicated. If the scattering of the particles is purely elastic, the one-particle analysis can be generalized in an obvious way: there is a function $\epsilon_a(\theta)$ for each particle type with mass $m_a$, and the TBA system (8) becomes a matrix equation involving \[ K_{ab}(\theta)= {1\over2\pi i}{d\over d\theta}\ln S_{ab}(\theta), \] where $S_{ab}(\theta)$ is the S-matrix element between particles $a$ and $b$. If the scattering is not elastic, there are still greater complications, with the net result being that the TBA system involves additional ``magnon'' degrees of freedom which behave like particles with zero mass. Nevertheless the analysis can sometimes be carried out successfully in this situation too, as we shall see later for a particular example. \section{SUMMARY OF RESULTS} We now summarize the results obtained for several families of integrable models. In each case we shall define the theory by a Lagrangian $\cal L$ displaying a global symmetry group $G$ or $G{\times}G$, where $G$ = SU($n$), SO($n$) or Sp($n$). To express the results compactly, we introduce the quantity $1/\Delta$ = $n$, $n{-}2$ or $2n{+}2$ respectively (this is the dual Coxeter number for SU($n$) and SO($n$) but {\it twice\/} the dual Coxeter number for Sp($n$)). The fields in the Lagrangian transform in the defining representation of the symmetry group unless we state otherwise. The mass-gap $m/\Lambda$ will be given for particles which also belong to the defining representation of the symmetry group. The renormalization scheme is $\overline{\rm MS}$ except in one case. {\bf (i)} O($n$) sigma model \cite{HN,HMN}: \[ {\cal L}={1\over2g}\, \partial_\mu \phi_a\partial^\mu \phi_a \] where $\phi_a$ is an $n$-component real scalar field obeying $\phi_a\phi_a=1$. The TBA calculation confirms the S-matrix proposed in \cite{ZZ}, correctly predicting $\beta_1 = 1/ 2\pi \Delta$ and $\beta_2 = 1/ 4\pi^2 \Delta$. The mass gap \[ m={(8 /e )^\Delta \over\Gamma(1+\Delta)} \, \Lambda_{\overline{\rm MS}}\ \] is consistent with the $1/n$ expansion \cite{BCR}. {\bf (ii)} $G{\times}G$ principal chiral model \cite{BNNW,THIII}: \[ {\cal L}={1\over g} \, {\rm Tr}\left(\partial_\mu U \partial^\mu U^{-1}\right) \] where $U$ is a $G$-valued field. The TBA calculation with the S-matrices proposed in \cite{SMPCM} correctly predicts $\beta_1 = 1/ 16 \pi x \Delta$, $\beta_2 = \beta_1^2 /2$ where $x$ is the Dynkin index of the defining representation of $G$ ($x=1/2$ for SU($n$), Sp($n$); $x=1$ for SO($n$)). The mass gap is \[ m= 2^{(d \Delta + 1 /2)} \, {\sin(\pi\Delta)\over \sqrt{\pi e}\Delta} \, \Lambda_{\overline{\rm MS}} \] where $d$ is the dimension of the defining representation of $G$ ($d=n$ for SU($n$), SO($n$); $d=2n$ for Sp($n$)). The mass gap has been measured on the lattice for $G$ = SU($3$) and the agreement with the theoretical value is quite accurate \cite{HAS}. {\bf (iii)} O($n$) Gross-Neveu model $(n > 4)$ \cite{FNW}: \[ {\cal L}={i\over2}\bar\psi_a\gamma^\mu\partial_\mu\psi_a+{g\over8} \left(\bar\psi_a\psi_a\right)^2 \] where $\psi_a$ is an $n$-component Majorana spinor. The S-matrices proposed in \cite{ZZ,SMGN} are found to be consistent with the TBA/perturbation calculation which correctly predicts $\beta_1 = 1/2\pi \Delta$ and $\beta_2 = -1/ 4\pi^2 \Delta$. The mass gap is \[ m= { \, \, (2e)^\Delta \over\Gamma(1- \Delta)} \, \Lambda_{\overline{\rm MS}} \] which is consistent with the $1/n$ expansion \cite{FNWII}. Note that our coupling constant $g$ and mass-scale $\Lambda_{\overline{\rm MS}}$ differ slightly from those used in \cite{FNW,FNWII}. {\bf (iv)} SU($n$) chiral Gross-Neveu model \cite{CGN}: \[ {\cal L}={i\over2}\bar\psi_a\gamma^\mu\partial_\mu\psi_a+{g\over8} \left \{ \left(\bar\psi_a\psi_a\right)^2 - \left(\bar\psi_a\gamma_5\psi_a\right)^2 \right \} \] where $\psi_a$ is an $n$-component complex spinor. The S-matrix for the massive sector conjectured in \cite{SMCGN} is consistent with the TBA/perturbation calculation which correctly predicts $\beta_1 = 1/ \pi \Delta$ and $\beta_2 = -1 /\pi^2 \Delta$. The mass gap \[ m= {\, \, (e/4)^{\Delta /2} \over \Gamma(1- \Delta)} \, \Lambda \] agrees with the $1/n$ expansion \cite{CGN}. See \cite{CGN} for full details of the renormalization scheme and note also our slightly different definitions of $g$ and $\Lambda$. {\bf (v)} O($n$) supersymmetric sigma model $(n>4)$ \cite{EH1}: \[ {\cal L}={1\over2g}\left\{(\partial_\mu \phi_a)^2 +i\bar\psi_a\gamma^\mu\partial_\mu\psi_a + {1\over4}\left(\bar\psi_a \psi_a\right)^2\right \} \] where the fields $\phi_a$ and $\psi_a$ are an $n$-component real scalar and spinor satisfying the constraints $\phi_a\phi_a=1$ and $\phi_a\psi_a=0$. The model is invariant under $N=1$ supersymmetry transformations mixing the bosons and fermions. The S-matrix conjectured in \cite{SW} is consistent with the TBA/perturbation calculation which correctly predicts $\beta_1 = 1/2 \pi \Delta$ and $\beta_2 = 0$. The mass gap \[ m=2^{2 \Delta} \, {\sin ( \pi\Delta) \over \pi \Delta} \, \Lambda_{\overline{\rm MS}} \] is consistent with the $1/n$ expansion \cite{JAG}. {\bf (vi)} SU($n$) or ${\bf C}P^{n-1}$ supersymmetric sigma model \cite{EH2}: \[ {\cal L}={1\over2g} \left \{ \, \vphantom{ {x \over x} } \left | (\partial_\mu{-}A_\mu) z_a \right |^2 + i \bar \psi_a \gamma^\mu (\partial_\mu{-}A_\mu) \psi_a + \, {1 \over 4} \left[ \left (\bar \psi_a \psi_a\right)^2 - \left(\bar \psi_a \gamma_5 \psi_a\right)^2 - \left(\bar \psi_a \gamma_\mu \psi_a\right )^2\right]\right\} \] where $A_\mu = \frac{1}{2}( z^*_a \partial_\mu z_a - z_a \partial_\mu z^*_a)$ and the fields $z_a$ and $\psi_a$ are an $n$-component complex scalar and Dirac spinor satisfying the constraints $z_a^* z_a^{\phantom{*}} = 1$ and $z_a^* \psi_a^{\phantom{*}} = 0$. The Lagrangian has a local U(1) invariance under which all complex fields transform by a phase and $A_\mu$ transforms as a gauge field. The model is also invariant under $N=2$ supersymmetry. The S-matrix proposed in \cite{SMCPN} correctly predicts $\beta_1 = 1/ \pi \Delta$ and $\beta_2 = 0$. The formula for the mass-gap is \[ m= {\sin ( \pi \Delta ) \over \pi \Delta} \, \Lambda_{\overline{\rm MS}} \] which is consistent with the $1/n$ expansion \cite{MC} and with other non-perturbative approaches specific to $N=2$ supersymmetric theories \cite{CV}. \section{DETAILED EXAMPLE: THE SUPERSYMMETRIC ${\bf C}P^{n-1}$ MODEL} Most of the features of the general method we have outlined here are nicely illustrated by the last example, the supersymmetric ${\bf C}P^{n-1}$ model \cite{INST,DDL2}, so we now discuss this case in more detail. To introduce the S-matrix, we assume that in the quantum theory there exist states $\vert a,i,\theta\rangle$ representing fundamental particles and states $\vert\bar a,i,\theta\rangle$ representing fundamental antiparticles. Here $\theta$ is rapidity; $i=0,1$ distinguishs ``bosons'' and ``fermions'' (actually these particles carry fractional statistics \cite{SMCPN,FI}); and indices $a$ and $\bar a$ label the $n$ and $\bar n$ representations of $\SU$. The fundamental anti-particles can be regarded as bound-states of the fundamental particles, or vice-versa---an example of ``nuclear democracy''. There are additional bound states transforming in all the antisymmetric representations of SU($n$). The integrability of the model \cite{AAG} implies that the S-matrix factorizes and that all S-matrix elements can be deduced from the two-body ones; furthermore, the S-matrix for any desired set of particles can be obtained from the S-matrix for the fundamental particles. Thus the entire S-matrix is specified by the amplitude \begin{equation} \langle c,k,\theta_2;d,l,\theta_1,{\rm out}\vert a,i,\theta_1;b,j,\theta_2,{\rm in}\rangle =S_{N=2}(\theta_1-\theta_2)_{ij}^{kl} S_{\rm CGN}(\theta_1-\theta_2)_{ab}^{cd} \end{equation} which was first proposed by K\"oberle and Kurak \cite{SMCPN}. Following \cite{Schou,FI}, we have written this proposal in a factorized form in which $S_{\rm CGN}$ is the S-matrix for the chiral Gross-Neveu model \cite{SMCGN}, which specifies the scattering of the $\SU$ degrees of freedom, and $S_{N=2}$ controls the $N=2$ supersymmetric degrees of freedom, as described in \cite{FI} (although we have chosen to include the physical strip pole in the CGN part of the S-matrix). Explicit expressions for these factors are given in \cite{EH2}. The conjectured S-matrix is minimal in the sense that it has the minimum number of poles and zeros on the physical strip (the region $0\leq{\rm Im}(\theta)\leq\pi$) consistent with the requirements of symmetry, the existence of a bound-state and the axioms of S-matrix theory. But this still leaves open the possibility of adding CDD factors to the S-matrix \cite{CDD,ZZ}; these spoil none of the axioms, they introduce no new poles on the physical strip and they passively respect the bootstrap equations. For our model the CDD ambiguities correspond to multiplying the S-matrix of the fundamental particles by factors of the form \begin{equation}{ \sinh\left({\theta\over2}-{i\pi\over2 n}\alpha\right) \sinh\left({\theta\over2}-{i\pi\over2 n}(2-\alpha)\right) \over \sinh\left({\theta\over2}+{i\pi\over2 n}\alpha\right) \sinh\left({\theta\over2}+{i\pi\over2 n}(2-\alpha)\right) } \ , \end{equation} where $0<\alpha<2$. One of the conclusions of our analysis will be that the minimal form is the true S-matrix of the theory, so that all CDD factors are ruled out. We have emphasized that in following the general method for testing the equivalence between an S-matrix and a Lagrangian, there are at least two important points to be borne in mind when choosing the coupling to a conserved charge $Q$. First, in order to simplify the analysis of the TBA system we must choose $Q$ so that the new ground-state consists of a restricted number of particle types. In the models considered in \cite{HMN,HN,FNW,CGN,BNNW,THIII} it was possible to find a generator $Q$ such that the new ground-state contained a single particle type. In the present theory, however, we know that the lowest energy states must come in degenerate supersymmetric multiplets, since supersymmetry commutes with the $\SU$ invariance. We might be tempted to make the same choice for $Q$ as in the $\SU$ principle chiral model \cite{BNNW} and the chiral Gross-Neveu model \cite{CGN}: \begin{equation} Q={\rm diag}\left(1,-{1\over n-1},\ldots,-{1\over n-1}\right), \end{equation} for which there will be a single fundamental doublet $\vert1,j,\theta\rangle$ with the largest charge/mass ratio. But it turns out that this violates our second criterion regarding the choice of $Q$, which is that there should be a classical or tree-level term in the ground-state energy so that we can extract the mass-gap via a one-loop, rather than a higher-loop, calculation. We are therefore motivated to consider an alternative choice: \begin{equation} Q={\rm diag}\left(1,-1,0,\ldots,0\right). \end{equation} This does indeed lead to the desired tree-level term, but the TBA analysis is complicated by the fact that there are now two fundamental doublets with the largest charge/mass ratio, namely $\vert1,j,\theta\rangle$ and $\vert\bar2,j,\theta\rangle$, Despite this complication, the calculation proves tractable with this choice, as we shall describe below. \subsection{Perturbation theory calculation} The coupling of the theory to the charge (15) by a change in the Hamiltonian $H \rightarrow H - h Q$ can be achieved by making a corresponding replacement $\partial_0 \rightarrow \partial_0 + i h Q$ in the Lagrangian. Since we are interested in performing a one-loop calculation of the change in the ground-state energy density as a function of $h$, it is enough to expand the resulting Lagrangian to quadratic order in an independent set of fields, and we can drop all terms which are independent of $h$ to this order. By exploiting the local U(1) invariance of the Lagrangian, we can take $z_1$ to be real and we can solve the bosonic constraint $z_a z_a^* =1$ by writing $z_1 = \{ (1 - \vert \pi \vert^2 )(\frac{1}{2} + \phi) \}^{1/2}$ and $z_2 = e^{i \theta} \{ (1 - \vert \pi \vert^2)(\frac{1}{2} - \phi) \}^{1/2}$ where $\pi = (z_3 , \ldots , z_n)$ and $\theta$, $\phi$ are real. The fermionic degrees of freedom and the variable $\theta$ decouple to quadratic order and we are left with the expression \[ {\cal L}_{{\rm 1-loop}} = {1 \over 2 g} \left \{ (\partial_\mu \phi)^2 + \vert \partial_\mu \pi \vert^2 + h^2 - 4 h^2 \phi^2 - h^2 \vert \pi \vert^2 \right\}. \] We see that there is indeed a tree-level term, as desired. Using standard dimensional regularization with the $\overline{\rm MS}$-scheme gives a one-loop expression for the ground-state energy \[ \delta {\cal E}(h) = -{h^2 \over 2 g} - {h^2 \over \pi} \ln 2 + {h^2 n \over 4 \pi} \left [ 1 - \ln {h^2 \over \mu^2} \right ]. \] It is important that when we substitute for the running coupling with the known values of the beta-function coefficients \begin{equation} \beta_1 = n / \pi \ , \quad \beta_2 = 0 \ , \end{equation} the $\mu$-dependence must cancel (to leading order) since the quantity we are computing is renormalization group-invariant. {}From above, the values of the other dimensionless numbers appearing in (1) are, for this model, \begin{equation} \alpha_0 = -1/ 2 \ , \quad \alpha_1 = (n/4 \pi) - (1/\pi) \ln2 \ . \end{equation} The explicit expression for the ground-state energy is thus, to the required order, \begin{equation} \delta {\cal E}(h)=-{h^2n\over2\pi}\left[ \, \ln{h\over \Lambda_{\overline{\rm MS}}}-{1\over2}+{2 \over n} \ln 2 \, \right]. \end{equation} \subsection{TBA calculation} We must now consider the TBA equations for the model and solve them in the limit $h\gg m$. Following the hypothesis introduced earlier, we assume that only the multiplets $\vert1,j,\theta\rangle$ and $\vert\bar2,j,\theta\rangle$ contribute to the ground-state. Since the scattering of these multiplets is purely elastic, it is not necessary to perform a diagonalization in the space of $\SU$ quantum numbers (although this diagonalization can be done \cite{THII}). The remaining difficulty is that the S-matrix for these favoured states is still non-diagonal in the supersymmetric subspace. Fortunately, this problem can be solved---in fact it has been shown by Fendley and Intriligator \cite{FI} that it is equivalent to diagonalizing the transfer matrix of the six vertex model at the free fermion point. The TBA equations involve energy densities $\epsilon_1(\theta)$ and $\epsilon_{\bar2}(\theta)$ for the two supermultiplets and two magnon energy densities $\xi_0(\theta)$ and $\xi_{\bar 0}(\theta)$ which reflect the non-elastic nature of the scattering amongst the supersymmetric degrees of freedom (two magnons because we are dealing with $N=2$ supersymmetry). The resulting TBA equations are, at $T=0$, \begin{eqnarray} \epsilon_a(\theta)-\phi_{ab}*\epsilon^-_b(\theta)-\phi_{al}*\xi^-_l(\theta) \!\!\!&=&\!\!\!m\cosh\theta-h,\nonumber\\ \xi_l(\theta)-\phi_{al}*\epsilon^-_a(\theta) \!\!\!&=&\!\!\!0, \end{eqnarray} where $a=1,\bar2$ and $l=0,\bar0$. In terms of these the ground-state energy density is \begin{equation} \delta {\cal E}(h)={m\over2\pi}\int_{-\infty}^\infty d\theta \left[\epsilon_1^-(\theta)+\epsilon_{\bar2}^-(\theta)\right]\cosh\theta. \end{equation} The kernels appearing in (19) are \begin{eqnarray*} \phi_{ab}(\theta)\!\!\! &=&\!\!\!{1\over2\pi i}{d\over d\theta} \ln S_{\rm CGN}(\theta)_{ab}^{ba},\\ \phi_{10}(\theta)\!\!\! &=&\!\!\!\phi_{\bar2\bar0}(\theta)= {1\over2\pi}{\sin(\pi/n)\over\cosh\theta-\cos(\pi/n)},\\ \phi_{\bar20}(\theta)\!\!\! &=&\!\!\!\phi_{1\bar0}(\theta)= {1\over2\pi}{\sin(\pi/n)\over\cosh\theta+\cos(\pi/n)}, \end{eqnarray*} where $S_{\rm CGN}(\theta)_{ab}^{ba}$ are the appropriate S-matrix elements of the chiral Gross-Neveu model. To simplify (19) it is important to notice that $\phi_{al}(\theta)$ is a positive kernel, which implies that the magnon variables are given by $\xi^+_l(\theta)=0$ and $\xi^-_l(t)=\phi_{al}*\epsilon^-_a(\theta)$. Furthermore, the solution does not distinguish between the values of the favoured SU($n$) quantum numbers and so we have $\epsilon_1(\theta)=\epsilon_{\bar2}(\theta)\equiv \epsilon(\theta)$. The four equations in (19) then reduce to a single equation for $\epsilon(\theta)$ of the form (8) with kernel \[ R(\theta)=\delta(\theta)-\phi_{11}(\theta)-\phi_{1\bar2}(\theta) -\left[\phi_{10}+\phi_{1\bar0} \right]*\left[\phi_{10}+\phi_{1\bar0}\right](\theta) \ , \] and, from (20), each of the densities $\epsilon_1$ and $\epsilon_{\bar 2}$ contributes an amount to the ground-state energy given by (9). We have therefore succeeded in reducing the problem to the relatively simple case which we already know how to handle. Following the subsequent steps outlined in section 2, we find the Fourier transform of the kernel: \[ \hat R(\omega) ={\cosh(({1\over2}{-}{1\over n})\pi\omega)\sinh({1 \over n}\pi\omega) \over\cosh^2(\frac{1}{2}\pi\omega)} \, \exp{\textstyle \frac{1}{2}\pi\omega} \ . \] This vanishes at the origin and we can decompose it as $1/(G_+(\omega)G_-(\omega))$ where $G_\pm(\omega)$ are analytic in the upper/lower half planes and $G_-(\omega)=G_+(-\omega)$. The unique solution is \begin{eqnarray*} G_+(\omega)={\Gamma(\frac{1}{2}-(\frac{1}{2}{-}{1 \over n})i\omega) \Gamma(1-{1 \over n}i\omega)\over \Gamma^2(\frac{1}{2}-\frac{1}{2} i\omega)} \exp \left \{ \, \textstyle \frac{1}{2} \ln n -\frac{1}{2}(1{+}i\omega) \ln(-i\omega) \right. \\ \textstyle \left. + \, i \omega \left [ \, \ln 2 + \frac{1}{2} + (\frac{1}{2}{-}{1\over n}) \ln (\frac{1}{2}{-}{1\over n})+{1\over n}\ln{1\over n} \, \right ] \, \right \} \end{eqnarray*} and we can indeed find an expansion of $G_+(i\xi)$ for small $\xi$ of the form (10) with \[ k=\sqrt{n/\pi}\ , \quad a=-1/2\ , \quad b= {\textstyle \frac{1}{2} (1{-}\gamma_{\rm E})- {1\over n} \ln 4n + ( \frac{1}{2}{-}\frac{1}{n} ) \ln ( {1\over 2}{-}{1\over n} ) } \ . \] Substituting in (11) and adding the contributions for the densities $\epsilon_1$ and $\epsilon_{\bar 2}$ produces \begin{equation} \kappa_0 = -n/ 2 \pi \, , \quad \kappa_1 = 0 \, , \quad \kappa_2 / \kappa_0 = \ln (n / \pi) + \ln \sin (\pi / n) + (2 \ln 2)/n - 1/2 \, , \end{equation} or explicitly, to the required order, \begin{equation} \delta {\cal E}(h)=-{h^2n\over2\pi}\left[ \, \ln{h\over m} - \frac{1}{2} + {2 \over n} \ln 2 +\ln\left({n\over \pi}\right) +\ln \sin \left({\pi \over n}\right) \right] \ . \end{equation} \subsection{Comparison of calculations} By substituting the values (16), (17) and (21) into the general relations (3) and (4)---or simply by comparing the explicit final results (18) and (22)---we see that the TBA calculation correctly reproduces the universal coefficients of the beta-function and that it predicts the value of the mass-gap for the supersymmetric ${\bf C}P^{n-1}$ model to be $ m /\Lambda_{\overline{\rm MS}}=(n/\pi)\sin(\pi/n)$, as claimed. It is instructive to consider what would have happened if we had carried out our analysis using the other charge (14), rather than (15). In that case only the multiplet $\vert1,j,\theta\rangle$ appears in the ground-state and the resulting TBA equations are simpler in as much as they involve only this single doublet, rather than two doublets. The system can be reduced to a single integral equation in a similar way, but on doing so we find a different kernel: \[ R(\theta)=\delta(\theta)-\phi_{11}(\theta) -\phi_{10}*\phi_{10}(\theta)- \phi_{1\bar0}*\phi_{1\bar0}(\theta) \ . \] The Fourier transform of this kernel does not vanish at the origin, and so, as explained previously, we would need to go beyond one-loop perturbation theory to carry out a non-trivial test of the S-matrix. This matches precisely the fact that with this different choice of charge the perturbative expansion of the ground-state energy density is also markedly different with no tree-level contribution. It is also interesting to see for this particular example how the calculation resolves the problem of CDD ambiguities in the S-matrix. An additional CDD factor of the form (13) would alter the kernel appearing in the TBA equation from $\hat R(\omega)$ to \[ \hat R(\omega)-2{\cosh(({1\over2}{-}{1\over n})\pi\omega) \cosh({1\over n}(\alpha-1)\pi\omega)\over \cosh(\frac{1}{2}\pi\omega)} \ . \] But this expression fails to vanish at the origin and so the argreement with the perturbative result is destroyed.
\section{Introduction} The phenomenology of neutralino dark matter has been studied extensively in the Minimal Supersymmetric extension of the Standard Model (MSSM) \cite{Susy}. This model incorporates the same gauge group as the Standard Model and the supersymmetric extension of its particle content. The Higgs sector is slightly modified as compared to that of the Standard Model: the MSSM requires two Higgs doublets $H_1$ and $H_2$ in order to give mass both to down-- and up--type quarks and to cancel anomalies. After Electro--Weak Symmetry Breaking (EWSB), the physical Higgs fields consist of two charged particles and three neutral ones: two scalar fields ($h$ and $H$) and one pseudoscalar ($A$). The Higgs sector is specified at the tree level by two independent parameters: the mass of one of the physical Higgs fields and the ratio of the two vacuum expectation values, usually defined as $\tan\beta=v_2/v_1 \equiv <H_2> / <H_1>$. The supersymmetric sector of the model introduces some other free parameters: the mass parameters $M_1$, $M_2$ and $M_3$ for the supersymmetric partners of gauge fields (gauginos), the Higgs--mixing parameter $\mu$ and, in general, all the masses of the scalar partners of the fermions (sfermions). In the MSSM it is generally assumed that the gaugino masses are equal at the grand unification scale $M_{GUT}$: $M_i(M_{GUT})\equiv m_{1/2}$ and hence are related at lower scales by \begin{equation} M_1 : M_2 : M_3 = \alpha_1 : \alpha_2 : \alpha_3 \label{eq:GUTgaugino} \end{equation} \noindent where the $\alpha_i$ (i=1,2,3) are the coupling constants of the three Standard Model gauge groups. The neutralinos are mass--eigenstate linear superpositions of the two neutral gauginos ($\tilde \gamma$ and $\tilde Z$) and the two neutral higgsinos ($\tilde H_1$ and $\tilde H_2$) \begin{equation} \chi = a_1 \tilde \gamma + a_2 \tilde Z + a_3 \tilde H_1 + a_4 \tilde H_2\;. \end{equation} \noindent The neutralino sector depends, at the tree--level, on the following (low--energy) parameters: $M_1= (5/3) \tan^2 \theta_W M_2$, $M_2 \simeq 0.8\, m_{1/2}$, $\mu$ and $\tan\beta$. Neutralino properties are naturally discussed in the ($m_{1/2}$, $\mu$) plane, for a fixed value of $\tan\beta$. As an example, in Fig.1 the lines of constant mass for the lightest neutralino ($m_{\chi}$) and constant gaugino fractional weight ($P \equiv a_1^2 + a_2^2$) are plotted in the ($m_{1/2}$, $\mu$) plane for $\tan\beta=8$. We observe that the mass of the lightest neutralino increases from the bottom left to the top right, while the neutralino composition changes from higgsino dominance in the top--left region of the plane to gaugino dominance in the bottom--right. The regions forbidden by accelerator data are also displayed in Fig.1. The low--energy MSSM scheme is a purely phenomenological approach, whose basic idea is to impose as few model--dependent restrictions as possible. In this approach the lightest neutralino is a favourite candidate for cold dark matter. This scheme has been employed extensively in the analysis of the size and the relevance of various possible signals of neutralino dark matter: direct detection \cite{thdir,elflor,direct}, signals due to neutralino annihilation in celestial bodies, namely the Earth and the Sun \cite{bodies,ap3}, and signals from neutralino annihilation in the galactic halo \cite{halo}. The MSSM provides a useful framework in which neutralino phenomenology may be analysed without strong theoretical prejudices which could, {\it a posteriori}, turn out to be incorrect. This scheme is also frequently employed in analyses of the discovery potential of future accelerators \cite{Baer}. At a more fundamental level, it is natural to implement this phenomenological scheme within the supergravity framework \cite{sugra,bbo,diehl}. One attractive feature of the ensuing model is the connection between soft supersymmetry breaking and EWSB, which would then be induced radiatively. The essential elements of the model are described by a Yang--Mills Lagrangian, the superpotential, which contains all the Yukawa interactions between the standard and supersymmetric fields, and by the soft--breaking Lagrangian, which models the breaking of supersymmetry. Here we only recall the soft supersymmetry breaking terms \begin{eqnarray} &-{\cal L}_{soft}& = \displaystyle \sum_i m_i^2 |\phi_i|^2 \no \\ &+& \left\{\left[ A^{l}_{ab} h_{ab}^{l} \tilde L_a H_1 \tilde R_b + A^{d}_{ab} h_{ab}^{d} \tilde Q_a H_1 \tilde D_b + A^{u}_{ab} h_{ab}^{u} \tilde Q_a H_2 \tilde U_b +\mb{h.c.} \right] - B \mu H_1 H_2 + \mb{h.c.} \right\} \no \\ &+& \displaystyle \sum_i M_i (\lambda_i \lambda_i + \bar\lambda_i \bar\lambda_i) \label{eq:soft} \end{eqnarray} \noindent where the $\phi_i$ are the scalar fields, the $\lambda_i$ are the gaugino fields, $H_1$ and $H_2$ are the two Higgs fields, $\tilde Q$ and $\tilde L$ are the doublet squark and slepton fields, respectively, and $\tilde U$, $\tilde D$ and $\tilde R$ denote the $SU(2)$--singlet fields for the up--squarks, down--squarks and sleptons. In Eq.(\ref{eq:soft}), $m_i$ and $M_i$ are the mass parameters of the scalar and gaugino fields, respectively, and $A$ and $B$ denote trilinear and bilinear supersymmetry breaking parameters, respectively. The Yukawa interactions are described by the parameters $h$, which are related to the masses of the standard fermions by the usual expressions, {\em e.g.}, $m_t = h^t v_2$. The supergravity framework is usually implemented with a number of restrictive assumptions about unification at $M_{GUT}$: i) Unification of the gaugino masses: $M_i(M_{GUT}) \equiv m_{1/2}$, ii) Universality of the scalar masses with a common mass denoted by $m_0$: $m_i(M_{GUT})$ \hfill \break \indent \phantom{ii)\ } $ \equiv m_0$, iii) Universality of the trilinear scalar couplings: $A^{l}(M_{GUT}) = A^{d}(M_{GUT}) = A^{u}(M_{GUT})$ \hfill \break \indent \phantom{iii)\ } $\equiv A_0 m_0$. \hfill\break \noindent These conditions have strong consequences for low--energy supersymmetry phenomenology, and in particular for the properties of the neutralino as dark matter particle. Typically, the lightest neutralino is constrained to regions of gaugino dominance, that entail a large relic abundance (in wide regions of the parameter space $\Omega_{\chi} h^2$ exceeds the cosmological upper bound) and a small direct detection rate for neutralino dark matter. Indirect signals from the neutralino, such as high--energy neutrinos from the Earth and Sun, and the products of annihilation in the halo, are practically undetectable \cite{diehl}. The above assumptions, particularly ii) and iii), are not very solid, since universality may occur at a scale higher than $M_{GUT}$, {\em i.e.}, the Planck scale or string scale \cite{comm}, in which case renormalization above $M_{GUT}$ may weak universality in the $m_i$, {\em e.g.}, between scalars in $\underline{\bar{5}}$ and $\underline{10}$ representations of $SU(5)$ \cite{ehnt}. Moreover, in many string models the $m_i$'s are not universal even at the string scale. In a number of recent works \cite{OlPok,others}, deviations from some of the unification conditions have been considered. In particular, in Ref.\cite{OlPok} phenomenological consequences for neutralinos of a relaxation of assumption ii) have been analysed in the regime of large values of tan $\beta$. It has been shown that deviations from condition ii) may entail a changeover in neutralino composition from a gaugino--like state to a higgsino--like state (or at least to a higgsino--gaugino mixed state), with important consequences for neutralino phenomenology. In this paper, we first explore, over the full range of tan $\beta$, the various scenarios which may occur when condition ii) is relaxed, with an approach which is similar to the one adopted in the large--$\tb$ analysis of Ref.\cite{OlPok}. We then discuss in detail the ensuing consequences for neutralino dark matter, with particular emphasis for its direct detection. In the following, we first discuss which constraints can be applied to the parameters when specific physical requirements are imposed. In Sect.II, we summarize the conditions implied by radiative EWSB and define the type of departure from universality examined in this paper. Then, in Sect.III we establish some upper bounds on the supergravity parameters by requiring that radiative EWSB does not occur with excessive fine tuning among different terms. In Sect.IV we analyse in detail the constraints due to the requirement that EWSB takes place radiatively. Subsequently, in Sect.V cosmological constraints, derived from the evaluation of the neutralino relic abundance, are discussed. Other constraints, from experimental data on $b \rightarrow s \gamma $ processes and on the mass of the bottom quark $m_b$, are applied in Sect.VI. In Sect.VII the effects of these various constraints are first displayed in the ($m_{1/2}$, $m_0$) plane for fixed $\tb$ and $A_0$, and then shown in the ($m_{1/2}$, $\mu$) plane, which provides the most useful representation for discussing neutralino phenomenology. We recall some specific properties of the neutral Higgs bosons in Sect.VIII. Finally, in Sect.IX event rates for direct detection of neutralino dark matter are discussed. Conclusions are presented in the last Section. \section{Radiative EWSB} We recall that the tree--level scalar potential for the neutral Higgs fields may be written in the form \begin{equation} V_0 = (M_{H_1}^2+\mu^2) |H_1|^2 + (M_{H_2}^2+\mu^2) |H_2|^2 -B\mu (H_1 H_2 + \mb{h.c.}) + \mb{quartic D terms.} \label{eq:higgspot} \end{equation} \noindent The parameters of this potential must obey the following physical conditions: \begin{equation} \sin 2 \beta = \frac {-2B\mu} {M_{H_1}^2+M_{H_2}^2+2 \mu^2} \label{eq:s2beta} \end{equation} \begin{equation} M_Z^2 = 2 \frac {M_{H_1}^2-M_{H_2}^2 \tan^2 \beta} {\tan^2 \beta -1} - 2 \mu^2 \label{eq:mz} \end{equation} \begin{equation} M_A^2 = M_{H_1}^2+M_{H_2}^2+2 \mu^2 > 0~. \label{eq:pinco} \end{equation} \noindent Here $M_A$ is the mass of the CP--odd neutral Higgs boson (see Sect.VIII below), and eq.(\ref{eq:pinco}) must in fact be strengthened to $M_A \geq (M_A)_{lb}$, where $(M_A)_{lb}$ is the experimental lower bound \cite{LEP}. For instance, for $\tb \gsim 3$, $(M_A)_{lb} \simeq 45$ GeV. Notice that the sign of $\mu$ is defined according to the convention of reference \cite{Susy}. We remark that although Eqs.(\ref{eq:higgspot}--\ref{eq:pinco}) are expressed at the tree level, in our actual calculations 1--loop corrections to $V_0$ \cite{oneloop} have been included. The $M_{H_i}$'s (as well as the sfermion and the gaugino masses and the parameters {\it A, B} and $\mu$) evolve from the $M_{GUT}$ scale down to the $M_Z$ scale according to the Renormalization Group Equations (RGE's). This is how Eq.(\ref{eq:mz}) may be satisfied, even if $M_{H_1}$ and $M_{H_2}$ are equal at $M_{GUT}$. In this work we consider deviations from universality in the scalar masses at $M_{GUT}$, which split $M_{H_1}$ from $M_{H_2}$. This effect is parameterized as \begin{equation} M^2_{H_i}(M_{GUT}) = m_0^2(1+\delta_i)~. \label{eq:nonuniv} \end{equation} \noindent The parameters $\delta_i$ which quantify the departure from universality for the $M^2_{H_i}$ will be varied in the range ($-1$,$+1$), but are taken to be independent of the supersymmetry parameters. This is an {\it Ansatz}, since, when evolving the scalar masses from the unification scale (Planck scale or string scale) to the GUT scale $M_{GUT}$, the deviation parameters are in general functions of all the supersymmetry parameters \cite{mark}. Following a common procedure, Eq.(\ref{eq:s2beta}) is used to replace the parameter $B$ by $\tan \beta$. Thus the set of independent parameters becomes $m_{1/2}$, $m_0$, $A_0$, $\tan \beta$, and $\mu^2$ is given in terms of these parameters by Eq.(\ref{eq:mz}), suitably corrected by 1--loop effects: only the sign of $\mu$ remains undetermined. Obviously, values of $\mu^2$ are accepted only if they exceed the experimental lower bound $\mu_{lb}^2$, which is derived from the lower limit on the chargino mass \cite{LEP}: $|\mu_{lb}| \simeq 45$ GeV. We have solved the RGE's using the 1--loop beta functions including the whole supersymmetric particle spectrum from the GUT scale down to $M_Z$, neglecting the possible effects of intermediate thresholds. Two--loop and threshold effects on the running of the gauge and Yukawa couplings are known not to exceed 10\% of the final result \cite{lang}. While this is of crucial importance as far as gauge coupling unification is concerned \cite{lang}, it is a second--order effect on the evolution of the soft masses. Since neutralino properties are studied over a wide range of variation for the high scale parameters, such a degree of refinement is not required here. In order to specify the supersymmetry phenomenology, boundary conditions for the gauge and Yukawa couplings have to be specified. Low--scale values for the gauge couplings and for the top--quark and the tau--lepton Yukawa couplings are fixed using present experimental results. In particular, we assign for the top mass the value $m_t=178$ GeV \cite{CDF}. In addition, we require the unification of the bottom and tau Yukawa couplings at the GUT scale, as would be suggested by a unifying group that includes an $SU(5)$--like structure \cite{chano}. The values of $M_{H_1}$ and $M_{H_2}$ at the $M_Z$ scale, obtained from the RGE's, may be parameterized in the following way: \begin{equation} M^2_{H_i} = a_i m_{1/2}^2 +b_i m_0^2 + c_i A_0^2 m_0^2+ d_i A_0 m_0 m_{1/2}~. \label{eq:polinom} \end{equation} \noindent (Notice that, in our notation, all running quantities written without any further specification are meant to denote their values at $M_Z$.) The coefficients in the expression (\ref{eq:polinom}) are functions of tan $\beta$ and of the $\delta_i$'s. They are displayed in Fig.2 (a,b) for the case of universal scalar masses, ({\em i.e.}, $\delta_i = 0$). The coefficients for $M_{H_2}^2$ turn out to be very stable as functions of $\tb$, except for small $\tb$. More precisely, $a_2 \sim -2.5$ for $\tb \gsim 4$ with all the other coefficients much smaller (of order 0.1). As far as $M_{H_1}^2$ is concerned, whereas $c_1$ and $d_1$ are again very stable (of order 0.1), $a_1$ and $b_1$ vary rapidly as functions of $\tb$. This property of $a_1$ and $b_1$ is due to the very fast increase of $h^b$ for increasing $\tb$. When a departure from $m_0$ universality is introduced, the coefficients in Eq.(\ref{eq:polinom}), except for $a_1$ and $a_2$, become functions of the parameters $\delta_i$: $b_1$, $c_1$ and $d_1$ depend on $\delta_1$ and $b_2$, $c_2$ and $d_2$ on $\delta_2$. Whereas the $b_i$'s are rapidly--increasing functions of the $\delta_i$'s, the $c_i$'s and the $d_i$'s are rather insensitive to these parameters. Stringent constraints on the parameters $m_{1/2}, m_0, A_0$ and $\tan \beta$ follow from the request that the ${M^2_{H_i}}$'s, evaluated from Eq.(\ref{eq:polinom}), satisfy Eqs.(\ref{eq:mz}--\ref{eq:pinco}). Explicitly, we require that $\mu^2$ and $M_A^2$, given by the expressions \begin{eqnarray} \mu^2 &=& \frac {1} { \tan^2 \beta -1} \{(a_1 - a_2 \tan^2 \beta) m_{1/2}^2+ (b_1 - b_2 \tan^2 \beta) m_0^2 + \nonumber \\ & & (c_1 - c_2 \tan^2 \beta) A_0^2 m_0^2 + (d_1 - d_2 \tan^2 \beta) A_0 m_0 m_{1/2}\}- \frac {M_Z^2} {2} \nonumber \\ &\equiv& J_1 m_{1/2}^2 + J_2 m_0^2 + J_3 A_0^2 m_0^2 + J_4 A_0 m_0 m_{1/2} - \frac {M_Z^2} {2} \label{eq:mu} \end{eqnarray} \begin{eqnarray} M_A^2 &=& (a_1 + a_2 +2 J_1) m_{1/2}^2+ (b_1 + b_2 +2 J_2 ) m_0^2 + \nonumber \\ & & (c_1 + c_2 +2 J_3) A_0^2 m_0^2 + (d_1 + d_2 +2 J_4) A_0 m_0 m_{1/2} - M_Z^2 \nonumber \\ &\equiv& K_1 m_{1/2}^2 + K_2 m_0^2 + K_3 A_0^2 m_0^2 + K_4 A_0 m_0 m_{1/2} - M_Z^2 \label{eq:M_A} \end{eqnarray} \noindent satisfy the conditions: $\mu^2 \geq \mu_{lb}^2$, $M_A \geq (M_A)_{lb}$ mentioned earlier. The coefficients $J_i$ and $K_i$ in Eqs.(\ref{eq:mu},\ref{eq:M_A}) are plotted as functions of $\tb$ in Fig.2 (c,d) for the case of $m_0$ universality. In Fig.2c we notice that all the $J_i$'s are positive, with $J_1$ dominating over the others: for $\tb \gsim 4$, one has $J_1 \simeq 2.4$. As far as the coefficients $K_i$ are concerned, we see in Fig.2d that only two of them are sizeable: $K_1$ and $K_2$. They are both decreasing functions of $\tb$, with $K_1 > K_2$. At very large $\tb$ these coefficients become very small, and $K_2$ even becomes negative (but still small in magnitude) at $\tb \gsim 50$. In the case of non--universality, the coefficients $J_i$ and $K_i$, except for $J_1$ and $K_1$, become functions of the parameters $\delta_i$. We will see in Sect.IV that many important features of the supersymmetry parameter space depend on the signs of the two coefficients $J_2$ and $K_2$. We show in Figs.3 and 4 how their signs depend on the values of the $\delta_i$'s. In Fig.3 the lines $J_2 = 0$ are plotted in the ($\delta_2$, $\delta_1$) plane for a few values of $\tb$: for each value of $\tb$, $J_2$ is negative in the region above the relevant $J_2 = 0$ line and positive below. Similarly, in Fig.4 the $K_2 = 0$ lines are displayed in the same ($\delta_2$, $\delta_1$) plane at fixed $\tb$: $K_2$ is negative above the $K_2 = 0$ lines, and positive below. We now make a few comments related to Eq.(\ref{eq:M_A}), since the value of $M_A$ plays a very crucial role in a number of important neutralino properties. This is due to the fact that many physical processes involving neutralinos are mediated by the neutral Higgs bosons. Thus the value of $M_A$ determines the size of the relevant cross sections both through $M_A^2$--dependence in propagators and, in an implicit way, through the couplings of the $h$ and $H$ bosons to quarks and to the lightest neutralino $\chi$ (see Sect.VIII). As a consequence, a small value of $M_A$ has the effect of enhancing the magnitude of the relevant cross sections. What values of $M_A$ do we obtain from Eq.(\ref{eq:M_A})? Because of the properties of the coefficients $K_i$ previously analysed, $M_A$ turns out to be a rapidly--decreasing function of $\tb$. In Fig.5, $M_A$ is displayed at the representative point $m_0 = 50$ GeV, $m_{1/2} = 200$ GeV (1--loop corrections to $M_A$ have been included in the calculation). One notices that $M_A$ is $O(M_Z)$ for $\tb \gsim 45$. This feature provides one of the most appealing scenarios for neutralino phenomenology. \section{Constraints due to the absence of fine tuning} Before we exploit fully the two constraints $\mu^2 \geq \mu_{lb}^2$, $M_A \geq (M_A)_{lb}$ to restrict the parameter space, we apply the general criterion that the expression (\ref{eq:mu}) is satisfied without excessive tuning among the various terms \cite{nft,bbo}. In radiative EWSB the physical value of $M_Z$, which sets the EW scale, may be written as \begin{eqnarray} M_Z^2 = 2( J_1 m_{1/2}^2 + J_2 m_0^2 + J_3 A_0^2 m_0^2 + J_4 A_0 m_0 m_{1/2} - \mu^2)~~. \label{eq:mz2} \end{eqnarray} Accidental compensation (fine tuning) among different terms in Eq.(\ref{eq:mz2}) may occur. We explicitly require the absence of too--strong fine tuning, {\em i.e.}, cancellations among exceedingly large values of the parameters $m_{1/2}$, $m_0$, $A_0$ and $\mu$. Denoting by $\eta_f$ a parameter which quantifies the degree of fine tuning, we require \cite{nft} that \begin{eqnarray} \left |\frac {\Delta M^2_Z} {M^2_Z}\right| & < & \eta_f \left |\frac {\Delta x^2_i} {x^2_i}\right | \label{eq:nft} \end{eqnarray} \noindent where $x_i$ denotes any of the previous parameters. For instance, for $A_0 = 0$, Eq.(\ref{eq:nft}) provides the following conditions \begin{eqnarray} m_{1/2}^2 < \frac {\eta_f} {2 |J_1|} M^2_Z, ~~ m^2_0 < \frac {\eta_f} {2 |J_2|} M^2_Z, ~~ \mu^2 < \frac {\eta_f} {2} M^2_Z \simeq (640\mb{ GeV})^2 \label{eq:pippo} \end{eqnarray} \noindent where in the last approximate equality we have taken $\eta_f = 100$, which means that we allow an accidental compensation at the 1\% level. The upper bound on $m_0$ depends on $\tb$ and the $\delta_i$'s, whereas that on $m_{1/2}$ varies only with $\tb$ (because of the nature of the {\it Ansatz} (\ref{eq:nonuniv}): see the comment after Eq.(\ref{eq:nonuniv})). For the sake of illustration, we give some numerical examples, taking again $\eta_f = 100$. For $\tb = 8$, we have, for $\delta_1 = \delta_2 = 0$, $m_{1/2} \lsim 400$ GeV, $m_0 \lsim 1.5$ TeV. For two other pairs of values of the $\delta_i$'s, which will be discussed later on, we obtain $m_0 \lsim 2.4$ TeV for $\delta_1 = -0.2$, $\delta_2 = 0.4$ and $m_0 \lsim 3.0$ TeV for $\delta_1 = -0.8$, $\delta_2 = 0.2$. At $\tb = 53$ we have $m_{1/2} \lsim 415$ GeV and $m_{0} \lsim (1.7-1.9)$ TeV, depending on the values for the $\delta_i$'s. These inequalities imply for the neutralino mass $m_\chi \lsim 170$ GeV. In the following, when graphical representations for the parameter space are shown, we display no--fine--tuning upper bounds obtained from the general expression (\ref{eq:nft}) with $\eta_f = 100$. These upper bounds are denoted by dashed lines in Figs.9--14. \section{Constraints due to radiative EWSB} The EWSB constraints are given by the set of Eqs.(\ref{eq:s2beta}--\ref{eq:pinco}), or equivalently by Eqs.(\ref{eq:mu}--\ref{eq:M_A}), together with the conditions $\mu^2 \geq \mu_{lb}^2$ and $M_A \geq (M_A)_{lb}$. From these equations the values of $m_0$ and $m_{1/2}$ (or $\mu$ and $m_{1/2}$) are constrained and thus some domains in the ($m_{1/2}$, $m_0$) or ($m_{1/2}$, $\mu$) planes can be excluded. Let us start this discussion by analyzing the condition $M_A \geq (M_A)_{lb}$, with $M_A$ given by Eq.(\ref{eq:M_A}). For the sake of simplicity, we put $A_0 = 0$ for the moment. To discuss the role of $M_A \geq (M_A)_{lb}$ in placing bounds on $m_{1/2}$ and $m_0$, we first rewrite it explicitly as \begin{eqnarray} K_1 m_{1/2}^2 + K_2 m_0^2 \geq M_Z^2 + (M_A)_{lb}^2~. \label{eq:con1} \end{eqnarray} \noindent The nature of this quadratic form in the ($m_{1/2}$, $m_0$) plane obviously depends on the signs of the two coefficients $K_1$ and $K_2$. As we have seen in Sect.II, it turns out that, whereas $K_1$ is always positive, the sign of $K_2$ depends on the values of $\tb$ and of the $\delta_i$'s. Two different situations may occur, depending on the sign of $K_2$. In the case $K_2 > 0$ the region allowed by (\ref{eq:con1}) is the one above an elliptical branch centered in the origin of the ($m_{1/2}$, $m_0$) plane. Therefore, both parameters $m_{1/2}$ and $m_0$ are bounded from below. When $K_2 < 0$, the region allowed by Eq.(\ref{eq:con1}) is the one between the $m_{1/2}$ axis and an upward--moving hyperbolic branch. Thus, whereas $m_{1/2}$ is still bounded from below, $m_0$ is now constrained from above. The upper bound on $m_0$ is particularly stringent when $K_2$ is large and negative and $K_1$ is not large. This occurs, for instance, at very large values of $\tb$ in the case of $m_0$ universality. This discussion may be extended straightforwardly to the case $A_0 \neq 0$. In this case the constraint $M_A \geq (M_A)_{lb}$ may be written explicitly as \begin{eqnarray} K_1 m_{1/2}^2 + K_2 m_0^2 + K_3 A_0^2 m_0^2 + K_4 A_0 m_0 m_{1/2} \geq M_Z^2 + (M_A)_{lb}^2~. \label{eq:con2} \end{eqnarray} \noindent The nature of this quadratic form depends on the sign of its determinant. When this determinant is positive, an elliptical branch in the ($m_{1/2}$, $m_0$) plane provides lower bounds on the two variables. On the other hand, a negative determinant entails an upward--moving hyperbolic branch which places an upper bound on $m_0$. These branches are part of conics whose axes are somewhat tilted with respect to the ($m_{1/2}$, $m_0$) axes. Similar implications follow from the constraint $\mu^2 \geq \mu_{lb}^2$, which may be written explicitly as (for $A_0 = 0$) \begin{eqnarray} J_1 m_{1/2}^2 + J_2 m_0^2 \geq \frac {M_Z^2} {2} + \mu_{lb}^2~. \label{eq:con3} \end{eqnarray} \noindent This quadratic form may be discussed in much the same way as the one in Eq.(\ref{eq:con1}). {}From the properties seen in Sect.II it turns out that the coefficient $J_1$ is always positive, whereas the coefficient $J_2$ is positive in the universal case, but may be negative when deviations from $m_0$ universality are introduced. Thus it follows that the condition $\mu^2 \geq \mu_{lb}^2$ puts lower bounds on $m_{1/2}$ and either lower or upper bounds on $m_0$, depending on the sign of $J_2$ (due to analytic properties identical to those discussed previously below Eq.(\ref{eq:con1})). The condition $\mu^2 \geq \mu_{lb}^2$ sets a very stringent upper bound on $m_0$, whenever $J_2$ is negative and large in magnitude. The extension to the case $A_0 \neq 0$ may be repeated here in a way similar to the above discussion for Eq.(\ref{eq:con2}). Thus we have seen that two important constraints, $\mu^2 \geq \mu_{lb}^2$ and $M_A \geq (M_A)_{lb}$, are at work in bounding $m_{1/2}$ and $m_0$, when EWSB is required to occur radiatively. When $J_2$ and $K_2$ are positive, the two conditions place lower bounds on $m_{1/2}$ and $m_0$. Similar constraints are established by the requirements that also the sfermion masses and $m_{\chi}$ satisfy the relevant experimental bounds. These last conditions are not explicitly discussed here, but they are taken into account in our evaluations. It is worth emphasizing that the most dramatic impact of the conditions $\mu^2 \geq \mu_{lb}^2$ and $M_A \geq (M_A)_{lb}$ over the parameter space occurs when either $J_2$ or $K_2$ (or both of them) are negative. Under these circumstances, as we have seen above, $\mu^2 \geq \mu_{lb}^2$ and $M_A \geq (M_A)_{lb}$ may place stringent upper limits on $m_0$, bounding the neutralino parameter space considerably. Which of the two conditions prevails over the other depends on the specific regions of the full parameter space and on the values of the $\delta_i$'s. In Sect.VII we will illustrate the implications of these constraints in a few specific examples. \section{Cosmological constraint} Let us turn now to the evaluation of the neutralino relic abundance $\Omega_{\chi}h^2$ and to the requirement that the lightest neutralino is not overproduced, {\em i.e.}, $\Omega_{\chi}h^2 \leq 1$. The neutralino relic abundance $\Omega_{\chi} h^2$ is evaluated following the standard procedure \cite{omega,omega_poles,omega1,ap1}, according to which $\Omega_{\chi} h^2$ is essentially given by $\Omega_{\chi} h^2 \propto <\sigma_{\mbi{ann}} v>^{-1}_{\mbi{int}}$, where $<\sigma_{\mbi{ann}} v>_{\mbi{int}}$ is the thermally--averaged annihilation cross section, integrated from the freeze--out temperature to the present temperature. The standard expansion $<\sigma_{\mbi{ann}} v> = a + b x + ...$ may be employed, with $x=T/m_\chi$, except at s--channel resonances ($Z,A,H,h$), where a more precise treatment has to be used for the thermal average \cite{omega_poles}. In the evaluation of $<\sigma_{\mbi{ann}} v>$ the full set of annihilation final states ($f \bar{f}$ pairs, gauge--boson pairs, Higgs--boson pairs and Higgs--gauge boson pairs), as well as the complete set of Born diagrams are taken into account \cite{ap1}. We recall that one of the largest contributions to the annihilation cross section is provided by diagrams with the exchange of the pseudoscalar Higgs boson $A$. (More relevant properties of the Higgs bosons are discussed in Section VIII.) We note that the constraint $\Omega_{\chi} h^2 \leq 1$ is very effective for small and intermediate values of $\tan \beta$, but is not restrictive for large values of $\tan \beta$. The strong restriction in the former case comes from the large value of $M_A$ implied by small and intermediate values of $\tan \beta$ (see Fig.5) (also the couplings of $A$ to $\chi$ and fermions are small for these values of $\tan \beta$). We show in Figs.6--8 a few examples where $\Omega_{\chi} h^2$ is given as a function of $m_{\chi}$ in the form of scatter plots. These scatter plots have been obtained by varying the parameters $m_0$ and $m_{1/2}$ on a equally--spaced linear grid over the ranges $10~\mb{GeV} \leq m_0 \leq 2~\mb{TeV}$, $45~\mb{GeV} \leq m_{1/2} \leq 500~\mb{GeV}$. Furthermore, we remark that all evaluations presented in this paper are for positive values of $\mu$, since negative values of $\mu$ are disfavoured by the constraints due to $m_b$ and $b \rightarrow s \gamma$ processes (see Sect.VI). The configurations shown in Figs.6--8 satisfy the constraints due to radiative EWSB, discussed previously. In Fig.6 is shown the case $\tb = 8$ and $\delta_i=0$. Here, as expected because of the intermediate value of $\tb$, many neutralino configurations provide $\Omega_{\chi}h^2 > 1$, whilst only a few give $\Omega_{\chi} h^2 \leq 1$. (Also, $M_A$ is large here because of sizeable values of $K_2$ (see Fig.2d), which helps increase $\Omega_{\chi}h^2$.) An exception occurs when $m_\chi \simeq M_Z/2$, since in this case the annihilation cross section is greatly enhanced due to the Z--pole contribution. In Fig.7 we display $\Omega_{\chi} h^2$ in a case of non--universality ($\delta_1 = -0.2, \delta_2 = 0.4$, for definiteness). It is easier to find $\Omega_{\chi}h^2 \leq 1$ in this case, since here the departure from $m_0$ universality implies a changeover of the neutralino composition from the gaugino dominance of the previous example to higgsino dominance (this point will be elucidated in Sect.VII). This implies a larger $\chi$--$\chi$ annihilation cross section and consequently a smaller relic abundance. Thus only a few neutralino configurations are excluded by the $\Omega_{\chi} h^2 \leq 1$ condition. An example for $\Omega_{\chi} h^2$ in the case of large $\tan \beta$ and $\delta_i = 0$ is shown in Fig.8. We see that $\Omega_{\chi} h^2 \leq 1$ imposes no constraint since, for this very large value of $\tb$, annihilation cross sections are very large. \section{Constraints from $\lowercase{b} \rightarrow \lowercase{s} \gamma$ and $\lowercase{m_b}$} In the evaluation of the $b \rightarrow s \gamma$ decay rate we have included the supersymmetric contributions arising from the charged Higgs loops and chargino loops given in Ref.\cite{bsgamma1}. The Higgs term always adds to the Standard Model value and usually entails too large a value for the rate. On the other hand, the chargino contribution gives rise to a destructive interference for $\mu > 0$ (in our convention for the sign of $\mu$). At large $\tan\beta$ supersymmetric contributions may be sizeable: unless the destructive interference protects the decay rate, it can very easily be driven out of the present experimental bounds. In the light of this property, the positive $\mu$ scenario appears to be the favourite one and, as already remarked, in this paper we only show results for this case. In comparing our predictions with observations we have taken into account that, as discussed in Ref.\cite{bsgamma_uncert}, large theoretical uncertainties are present, mainly due to QCD effects. In particular, predictions depend very strongly on the choice of the renormalization scale, leading to an inaccuracy of order 25\%. To account for this effect we have relaxed the experimental bounds of Ref.\cite{bsgamma_lim} by the same amount, keeping the renormalization scale fixed at the representative value of 5 GeV. Thus, our requirement is that the rate of $b \rightarrow s \gamma$ decay falls into the range $0.8 \times 10^{-4} \leq \mbox{BR}(b \rightarrow s \gamma) \leq 5.3 \times 10^{-4}~.$ The supersymmetric corrections to the bottom mass include contributions from bottom--squark--gluino loops and from top--squark--chargino loops \cite{carena}. In the present analysis, the bottom mass is computed as a function of the other parameters and required to be compatible with the present experimental bounds. Theoretical uncertainties in the evaluations of $m_b$ arise both from the running of the RGE's and from assumptions about Yukawa unification. Since our choice is to solve RGE's at the 1--loop level and without thresholds, we estimate an uncertainty of the order of 10\% in our prediction for $m_b$. In addition, a relatively small departure (see Ref.\cite{barger}) from bottom--$\tau$ Yukawa unification at the GUT scale may significantly change the bottom mass result. To take into account such uncertainties we have chosen to weaken the bounds on $m_b$ given in \cite{Wright} by an amount of 10\%. Thus we require $m_b$ to fall into the range $2.7~ \mbox{GeV} \leq m_b(M_Z) \leq 3.4~ \mbox{GeV}$. \section{Allowed regions in neutralino parameter space} We discuss now in a few examples how the various constraints analysed in the previous Sections complement each other in shaping the allowed regions in the parameter space. We start with the ($m_{1/2}$, $m_0$) representation, and later display our results in the ($m_{1/2}$, $\mu$) plane which provides the most useful representation for neutralino phenomenology. Let us first clarify a few graphical conventions adopted in our ($m_{1/2}$, $m_0$) and ($m_{1/2}$, $\mu$) plots. Regions are left empty when at least one of the following constraints is not satisfied: i) experimental bounds on Higgs, neutralino and sfermion masses \cite{LEP,d0}, ii) the $\chi$ is the Lightest Supersymmetric Particle (LSP), iii) radiative EWSB and $\mu^2 \geq \mu_{lb}^2$, $M_A \geq (M_A)_{lb}$. Regions forbidden by the cosmological constraint ($\Omega_{\chi} h^2 \leq 1$) are explicitly denoted by dots and those disallowed by the $b \rightarrow s \gamma$, $m_b$ constraints (but not by the previous ones) are denoted by crosses (crosses are displayed only in the ($m_{1/2}$, $m_0$) plane, but not in the ($m_{1/2}$, $\mu$) plane, to simplify these plots). The allowed domains are denoted by squares when they satisfy $\Omega_{\chi} h^2 > 0.01$, or by diamonds otherwise in the ($m_{1/2}$, $m_0$) plots. They are denoted by squares in the ($m_{1/2}$, $\mu$) plots, independently of the $\Omega_\chi h^2$ value. To simplify the discussion, we first take $A_0 = 0$. We comment on the $A_0 \neq 0$ case at the end of this Section. As a first example, let us consider the representative point $\tb =8$. For this intermediate value of $\tb$, the cosmological constraint is expected to be very effective in view of the arguments discussed in Sect.V. This is actually the case for universal $m_0$, when both $K_2$ and $J_2$ are positive (see Fig.2), so that the conditions of radiative EWSB do not set any upper limit on $m_0$ (Fig.9a). The empty region in the lower part of these figures is forbidden by the experimental bound on $m_{\chi}$. As shown in this figure, in wide regions (denoted by dots) $\Omega_{\chi} h^2 > 1$. Thus the cosmological constraint places a very stringent upper bound on $m_0$ for $m_{1/2} \gsim 150$ GeV. However, for smaller values of $m_{1/2}$, an allowed horizontal region extends up to $m_0 \simeq 2$ TeV. In fact, along this strip, $m_{\chi} \simeq M_Z/2$ and then $\Omega_{\chi} h^2 \leq 1$ is satisfied (see the discussion in Sect.V). Moving away from the universal point towards a region where $J_2$ is negative, we expect $\mu^2 \geq \mu_{lb}^2$ to be effective in placing a stringent upper bound on $m_0$. This is actually the case in the example shown in Fig.10a, which refers to the representative point $\delta_1 = - 0.2$, $\delta_2 = 0.4$ ($J_2 = -0.07$). Here it is the bound $\mu^2 \geq \mu_{lb}^2$ which provides the most stringent constraint in disallowing the large (empty) domain on the right side. Nevertheless, $\Omega_{\chi} h^2 \leq 1$ is still effective in excluding an internal region that would otherwise be allowed (see the discussion below). Keeping $\tb = 8$, we complete our discussion by considering the representative point $\delta_1 = - 0.8,\delta_2 = 0.2$ shown in Fig.11a, which gives an example where $J_2$ is very small. The peculiarity of this example will become clear when we discuss the relevant situation in the ($m_{1/2}$, $\mu$) plane, to which we now turn. The shape and general properties of the physical region in the ($m_{1/2}$, $\mu$) plane are dictated by the constraints previously derived, and they are determined most notably by $J_2$. It is convenient to distinguish the two cases i) $J_2 > 0$ and ii) $J_2 <0$. For case (i) at fixed $m_{1/2}$, $\mu$ increases for increasing $m_0$ with the consequence that the allowed physical region extends to the right of the $m_0=0$ line in the ($m_{1/2}$, $\mu$) plane, allowing for the neutralino only a gaugino--dominated region. In the case (ii) ($J_2 <0$), starting from the $m_0=0$ line and increasing $m_0$ at fixed $m_{1/2}$, one moves to the left and then one may reach regions of sizeable higgsino--gaugino mixing or even of higgsino dominance. Case i) applies in particular to the case of $m_0$ universality ($\delta_i = 0$) for any value of $\tb$. This is clear from Fig.3, which shows that in the ($\delta_2$, $\delta_1$) plane the origin is below any $J_2 = 0$ line. An example of this situation is displayed in Fig.9b (for $\tb = 8$). However, as we have seen in Sect.II, when the assumption of $m_0$ universality is relaxed, then $J_2$, which in the universal case is positive and small, may very easily become negative and sizeable. In this case a changeover in neutralino composition from an originally gaugino--like state into a higgsino--like one occurs. This remarkable property, discussed in Ref.\cite{OlPok} for large $\tb$, is in fact valid over the whole range of $\tb$, if the degree of non--universality is increased for decreasing $\tb$. An example of case ii) ($J_2 < 0$) is shown in Fig.10b, where the allowed region extends widely into the higgsino region. It is instructive to compare Fig.9 with Fig.10. Looking at sections a) of these figures, we notice that changing the values of the $\delta_i$'s from the set $\delta_i = 0$ to the set $\delta_1 = -0.2$, $\delta_2 = 0.4$ relaxes substantially the cosmological constraint. Parts b) of these figures provide the explanations for this feature. In fact, whereas in the former case the neutralino is mainly a gaugino, in the latter case $\chi$ is higgsino--like or mixed. As we already remarked, this implies an increase of the $\chi$--$\chi$ annihilation cross section and a reduction of the relic abundance. The physical region also displays an extension to the right, in the example of Fig.11b, but here the effect is very tiny, due to a very small $J_2$ and to the severe upper bound on $m_0$ for $m_{1/2} \gsim 180$ GeV. This is the first case to show a very marked ($m_{1/2}$, $\mu$) correlation. Now we turn to the case of large $\tb$, where new features appear. First, the $M_A \geq (M_A)_{lb}$ condition is no longer protected by large values of $K_1$, and may become effective in restricting the parameter space. Secondly, the $m_b$ and $b \rightarrow s \gamma$ conditions are now rather stringent over large domains and not only occasionally relevant as in the smaller $\tb$ cases. Thirdly, the cosmological constraint is usually overwhelmed by the other conditions. In Figs.12a, 13a, 14a we have, for $\tb = 53$, the following sequence of examples. i) $\delta_1 = 0, \delta_2 = 0$ (Fig.12a): here $K_2 < 0$, $J_2 > 0$, and since $K_2$ is negative and sizeable in magnitude, the constraint $M_A \geq (M_A)_{lb}$ sets an extremely stringent upper bound on $m_0$ and thus forbids the wide (empty) region on the right. ii) $\delta_1 = 0, \delta_2 = -0.2$ (Fig.13a): here one still has $K_2 < 0$, $J_2 > 0$, but $|K_2|$ is smaller than in the previous case, so the constraint $M_A \geq (M_A)_{lb}$ is still very effective but less compelling than in the case (i). Also, the role of the $m_b$ and the $b \rightarrow s \gamma$ conditions is more significant here. iii) $\delta_1 = 0.7, \delta_2 = 0.4$ (Fig.14a): here $K_2 > 0$, $J_2 < 0$, $M_A \geq (M_A)_{lb}$ gives a lower bound on $m_{1/2}$ and the $\mu^2 \geq \mu_{lb}^2$ condition provides the frontier of the empty domain on the right. The ($m_{1/2}$, $\mu$) representations for large $\tan \beta$ and for the representative $\delta_i$ points discussed above are displayed in Figs.12b--14b. We start from the universal case of Fig.12b. Here we expect gaugino--dominated configurations. However, because the values of $m_0$ are strongly limited from above (see Fig.12a), we have the extremely correlated states shown in Fig.12b. In the case of Fig.13b one has $J_2>0$, and gaugino--dominated states occur. No strong ($m_{1/2}$, $\mu$) correlation shows up in this case. The opposite case, $J_2 < 0$, is shown in Fig.14b, where higgsino--dominated configurations appear. It is worth adding a few comments about the examples of Figs.11 and 12, where the physical regions in the ($m_{1/2}$, $\mu$) plane show a very pronounced correlation in the two variables. This feature occurs whenever $|J_2| m_0^2 \ll J_1 m_{1/2}^2$, {\em i.e.}, whenever $m_0$ is severely bounded from above and/or $|J_2|$ is very close to zero. As far as the values of $|J_2|/J_1$ are concerned, we notice that in the universal case (see Fig.2c), except for small values of $\tb$, $J_2/J_1 \simeq 0.04$ (in fact, for $\tb \gsim 4$, $J_1 \simeq - a_2 \simeq 2.5$, $J_2 \simeq - b_2 \simeq 0.1$). Thus for $\delta_i = 0$ a strong ($m_{1/2}$, $\mu$) correlation occurs whenever $m_0 \lsim O(m_{1/2})$. This happens in the example of Fig.12, where $m_0$ is severely bounded by the $M_A \geq (M_A)_{lb}$ condition, and in the case of Fig.11, where the correlation is enforced by a very small value of $J_2$: $J_2 = 0.06$. A ($m_{1/2}$, $\mu$) correlation is also exhibited in Fig.9b for the range $m_{1/2} \gsim 150$ GeV, where $m_0$ is bounded by the cosmological constraint. In general, we do not consider these physical regions with a strong ($m_{1/2}$, $\mu$) correlation as unnatural, since they are usually realized without much tuning. We recall that the size of the coefficients $J_1$ and $J_2$ is dictated by the RGE's with their intrinsic cancellations, and that one naturally has $J_1= O$(a few), $J_2= O(0.1-0.01)$. As we have seen, these properties, combined with severe upper bounds on $m_0$, are sufficient to generate the ($m_{1/2}$, $\mu$) correlation. We turn now to the $A_0 \neq 0$ case. First we recall that $A_0$ is constrained in the range $|A_0| \lsim 3$ from the absence of charge and color breaking \cite{a_0}. Thus, allowing $A_0 \neq 0$ does not change essentially the general picture previously discussed. The previous scenarios still occur, but at different points in the parameter space. Two specific comments are in order here: i) independently of its sign, $A_0$ disfavours the changeover from gaugino dominance to higgsino dominance in the neutralino composition, ii) a negative $A_0$ reduces the value of $M_A$ as compared to the $A_0=0$ case, and so either provides a light $A$ boson (and hence interesting phenomenology) or enforces a more stringent constraint on the parameter space. \medskip \section{Neutral Higgs Bosons} Neutralino direct detection, to be discussed in the next Section, is based on neutralino--nucleus scattering. In this process, exchanges of neutral Higgs bosons play a dominant role, provided the Higgs masses are not too heavy. It is convenient to recall here some relevant properties of the couplings of $\chi$ with matter via Higgs exchange. As was already mentioned in the Introduction, the two Higgs isodoublets $H_1$, $H_2$ yield 3 neutral Higgs mass eigenstates: one $\mb{CP}$--odd ($A$) state, whose mass $M_A$ is given by expression (\ref{eq:M_A}) and two $\mb{CP}$--even states (of masses $M_h$, $M_H$, $M_h < M_H$), which are obtained from $H_1^0$, $H_2^0$ by a rotation through an angle $\alpha$ \beqarr H &=& \cos\alpha\,H_1^0 + \sin\alpha\,H_2^0 \nonumber \\ h &=& -\sin\alpha\,H_1^0 + \cos\alpha\,H_2^0 ~. \eeqarr \noindent It is important to notice here that $\alpha$ depends very sensitively on $M_A$, being very close to zero for $\tb \gsim 4$ and rising very fast to $\pi/2$ for $M_A \lsim O(M_Z)$ (see Fig.15). The angle $\alpha$ plays a crucial role in determining the size of the neutral $h,H$--quark couplings. Here, as we are interested in $\chi$--nucleus scattering, we discuss explicitly only the couplings involving the CP--even states, since $h,H$ are dominant compared to $A$. The low--energy neutralino--quark effective Lagrangian generated by Higgs exchange may be written as follows \cite{[4]} \beq {\cal L}_{\mbi{eff}} = \sqrt2 G_F {m_Z \over m_{h,H}^2} F_{h,H} \sum_q k_q m_q \bar{\psi}_{\chi} \psi_{\chi} \bar{q} q~. \label{eq:el} \eeq \noindent Here $F_{h,H}$ is the ratio of the Higgs--neutralino coupling to the $SU(2)$ gauge coupling, which depends on the composition of $\chi$ \beqarr F_h &=& a_2 (a_3 \sin \alpha + a_4 \cos \alpha) \nonumber \\ F_H &=& a_2 (a_3 \cos \alpha - a_4 \sin \alpha) \eeqarr \noindent and the $k_q$ are given, for the up--type quarks and the down--type quarks respectively, by \begin{eqnarray} \,\qquad &\qquad H \qquad & \qquad \qquad \quad h \qquad \nonumber \\ k_u \qquad & \qquad \sin\alpha / \sin\beta \qquad & \qquad \phantom{-} \cos\alpha / \sin\beta \qquad \nonumber \\ k_d \qquad & \qquad \cos\alpha / \cos\beta \qquad & \qquad - \sin\alpha / \cos\beta~. \end{eqnarray} \noindent Note that, in general, since $\tb > 1$, the strength of the coupling to the down--type quarks is bigger than the one to the up--type quarks, and ${\cal L}_{\mbi{eff}}$ usually gets a sizeable contribution when the $h$ boson is exchanged ($h$ is lighter than $H$ and is therefore favored because of the propagator denominator in Eq.(\ref{eq:el})) and when $\alpha \simeq \pi/2$, {\em i.e.}, when $M_A \lsim O(M_Z)$. When this regime does not apply, the size of ${\cal L}_{\mbi{eff}}$ is much suppressed. The cross section for elastic neutralino--nucleus scattering which follows from the effective Lagrangian (\ref{eq:el}) will be given in Sect.IX.B. \section{Direct detection} Much experimental activity is under way in the direct search for neutralino dark matter and the perspectives for significant improvements in experimental sensitivities are encouraging \cite{mosca}. In this class of experiments, a relic neutralino would be detected by the amount of energy released by its elastic scattering off nuclei in an appropriate apparatus. A signature would be provided by a yearly modulation of the signal, whose observations would require high statistics and extremely good stability in the detector response. Here we evaluate the event rates for this process extending previous analyses to the non--universal $\delta_i \neq 0$ case. Various materials are being used in the current experiments and others are under investigation for future detectors. In this paper we analyse two of the most interesting materials: Ge (in its natural composition) \cite{[15],[14],klapdor} and $^{129}$Xe \cite{[17]}. \subsection{Differential rates} The nuclear recoil spectrum may be evaluated from the expression \beq \frac {dR}{dE_R}= \sum_i \frac {R_{0,i}} {<E_R^{max}>} F_i^2(E_R)I(E_R) \label{eq:dir1} \eeq \noindent where \beq R_{0,i}=N_T \frac {\rho_{\chi}} {m_\chi} \sigma_i <v>~. \label{eq:dir2} \eeq \noindent In Eqs.(\ref{eq:dir1})--(\ref{eq:dir2}) we use the following notations: the subscript $i$ refers to the two cases of coherent and spin--dependent effective interactions, $N_T$ is the number of the target nuclei per unit of mass, $\rho_\chi$ is the local neutralino matter density, and $E_R$ is the nuclear recoil energy given by $E_R={{m_{red}^2}}v^2(1-\cos \theta^*)/{m_N}$, where $\theta^*$ is the scattering angle in the neutralino--nucleus center--of--mass frame, $m_N$ is the nuclear mass, $m_{\rm red}$ is the neutralino--nucleus reduced mass and $v$ is the relative velocity. The maximum value of $E_R$ is $E_R^{max}={{2m_{red}^2}}v^2/{m_N}$. Returning to (\ref{eq:dir1}--\ref{eq:dir2}), $F(E_R)$ denotes the nuclear form factor, and $\sigma_i$ is the (coherent/spin--dependent) neutralino--nucleus cross section. The factor $I(E_R)$ is given by \beq I(E_R)=\frac {<v^2>} {<v>} \int_{v_{min}(E_R)}^{v_{max}} dv \frac {f(v)} {v} \label{eq:dir3} \eeq \noindent where $f(v)$ is the velocity distribution of neutralinos in the Galaxy, as measured in the Earth's rest frame, and $v_{\rm min}(E_R)$ is given by $v_{\rm min}(E_R)=({{m_N E_R}/({2m_{\rm red}^2})})^{1/2}$. The averages appearing in Eqs.(\ref{eq:dir1})--(\ref{eq:dir3}) denote averages over the velocity distribution in the Earth's rest frame. An explicit formula for $I(E_R)$ in the case of a Maxwellian velocity distribution may be found in Ref.\cite{direct}. The differential rates to be discussed below will be expressed in terms of the electron--equivalent energy $E_{ee}$ rather than in terms of $E_R$. These two variables are proportional: $E_{ee}=Q E_R$ where $Q$ is called the quenching factor: typical values of $Q$ will be discussed shortly. \subsection{Neutralino--nucleus elastic cross sections} The total cross sections for neutralino--nucleus elastic scattering have been evaluated following standard procedures \cite{elflor,direct,[4],[5],goodman}. Here we only summarize some of the main properties. Neutralino--quark scattering is described by amplitudes with Higgs--boson exchanges and $Z$--boson exchange in the t--channel, and by amplitudes with squark exchanges in the s-- and u--channels. The neutral Higgs bosons considered here are the two CP--even bosons: $h,H$ and the CP--odd one: $A$, whose couplings were previously discussed in Sect.VIII. The relevant properties for these amplitudes are: 1) Higgs--boson exchanges contribute a coherent cross section which vanishes only when there is no zino--higgsino mixture in the neutralino composition \cite{[4]}, 2) $Z$--boson exchange provides a spin--dependent cross section which receives contributions only from the higgsino components of $\chi$, 3) squark exchanges contribute a coherent cross section (due to zino--higgsino mixing) as well as a spin--dependent cross section (due mainly to the gaugino components of $\chi$)\cite{[5]}. As examples we recall here only the expressions for the coherent cross section due to the exchange of a Higgs boson ($h$ or $H$) and the spin--dependent one due to Z exchange. The former cross section is easily evaluated from the effective Lagrangian of Eq.(\ref{eq:el}) \cite{[4]} \beq \sigma_{\mbi{CH}} = \frac {8 G_F^2} {\pi} \frac {m_Z^2} {m_{h,H}^4} \alpha^2_{h,H} m_{red}^2 A^2 \eeq \noindent where $A$ is the nuclear mass number and $\alpha_{h,H}$ is given by \beq \alpha_{h,H} = F_{h,H} I~~~~,~~~~ I=\sum_q k_q m_q \langle N|\bar{q} q |N \rangle . \eeq \noindent The quantity $I$ may be expressed conveniently in terms of the $\pi N$ sigma--term $\sigma_{\pi N}$ and of a parameter $a$ which is related to the strange--quark content of the nucleon $y$ by \beq a = y (m_s/(m_u + m_d))~~~,~~~~~ y = 2 \frac {<N|\bar{s} s|N>} {<N|\bar{u} u+\bar{d} d|N>}~. \eeq One has \beq I \simeq k_u g_u+k_d g_d \eeq \noindent where \beqarr g_u &=& {4 \over {27}} \left( m_N + {{19}\over{8}} \sigma_{\pi N} - a \sigma_{\pi N} \right) \nonumber \\ g_d &=& {2 \over {27}} \left( m_N + {{23}\over{4}} \sigma_{\pi N} + \frac {25} {2} a \sigma_{\pi N} \right) ~. \eeqarr \noindent Unfortunately, the values of both the quantities $y$ and $\sigma_{\pi N}$ are somewhat uncertain. Here, for $y$ we use the central value of the most recent evaluation: $y = 0.33 \pm 0.09$, obtained from a lattice calculation \cite{liu}. For $\sigma_{\pi N}$, which is derived by phase--shift analysis and dispersion relation techniques from low--energy pion--nucleon scattering cross--sections \cite{chen,gass}, we employ the value of Ref.\cite{gass}: $\sigma_{\pi N} = 45$ MeV. We then find the results: $g_u =123$ MeV, $g_d =288$ MeV (we use $2 (m_s/(m_u + m_d)) = 29$ \cite{bj}). We note that these values further reinforce the role of the down--type quarks as compared to the up--type ones. We point out that the Higgs--nucleon couplings for nucleons bound in a nucleus may be renormalized by the nuclear medium. As a consequence, the strength of $I$ might in principle be reduced to some extent \cite{brown}. However, this effect is neglected here. Now let us turn to the spin--dependent cross section due to Z exchange. This may be cast into the usual form \cite{[5],goodman} \beq \sigma_{\mbi{SD}} = \frac {8 G_F^2} {\pi} (a_3^2-a_4^2)^2 m_{red}^2 (\sum_q T_{3L,q} \Delta q)^2 \lambda^2 J(J+1)~. \label{eq:sd} \eeq \noindent In this paper we use this formula for $^{73}$Ge (this isotope is present at the level of 7.8 \% in the natural composition of Ge) and to $^{129}$Xe. For these nuclei we employ the values of $\lambda^2$ obtained in the odd--group model \cite{elflor}, where only the odd nuclear species in odd--even nuclei are explicitly taken into account. The $\Delta q$'s in Eq.(\ref{eq:sd}) denote the fractions of the nucleon spin carried by the quarks $q$ in the nucleon of the odd species, and the $T_{3L,q}$'s stand for the third components of the quark weak isospin. The values for the $\Delta q$'s are taken from Ref.\cite{ellkar}. It is worth noticing that the event rates for neutralino direct detection with the materials considered here are largely dominated by coherent effects in most regions of the parameter space. In the small domains where spin--dependent effects dominate over the coherent ones the total rates are usually too small to allow detection. The experimental strategy of employing materials enriched in heavy isotopes of high spin is interesting for a search for hypothetical dark matter particles which interact with matter via substantial spin--dependent interactions. However, this approach does not appear to be very fruitful for neutralinos. One more ingredient which enters the event rate in Eq.(\ref{eq:dir1}) is the nuclear form factor, which depends sensitively on the nature of the effective interaction involved in the neutralino--nucleus scattering. For the coherent case, we simply employ the standard parameterization \cite{[12]} \beq F(E_R)=3 \frac {j_1(qr_0)} {qr_0} e^{-\frac {1} {2}} s^2 q^2 \label{eq:ff} \eeq \noindent where $q^2 \equiv\mid{\vec {q}}\mid^2=2m_NE_R$ is the squared three--momentum transfer, $s \simeq 1~ \rm fm$ is the thickness parameter for the nucleus surface, $r_0 = (r^2-5s^2)^{1/2}$, $r=1.2~A^{1/3}$ fm and $j_1(qr_0)$ is the spherical Bessel function of index 1. The form factor in Eq.(\ref{eq:ff}) introduces a substantial suppression in the recoil spectrum unless $q r_0 \ll 1$. A noticeable reduction in $dR/dE_R$ may already occur at threshold $E_R=E_R^{\rm th}=E_{\mb{ee}}^{\rm th}/Q$ when $r_0 \sqrt{2 m_N E_R^{\rm th}}$ is not small compared to unity. The actual occurrence of this feature depends on parameters of the detector material: nuclear radius, quenching factor, threshold energy $E_{\mb{ee}}^{\rm th}$. The values of these parameters for the nuclei considered in this paper are reported in Table I \cite{mosca,[14],klapdor,[17]}, and the values of $F^2(E_R^{\mb{th}})$ calculated from Eq.(\ref{eq:ff}) are given in this same Table. Since we consider in this paper mainly the value of the differential rate near threshold, $F^2(E_R^{\mb{th}})$ is the most relevant quantity. We see from the values in the Table that the reduction introduced by the form factor is moderate in Ge, but quite substantial in $^{129}$Xe. In general, for the spin--dependent case there are no analytic expressions for the form factors. However, numerical analyses have been performed for a number of nuclei. The general feature is that these form factors have a much milder dependence on $E_R$ as compared to the coherent ones, because only a few nucleons participate in the neutralino--nucleus scattering in this case. In our evaluations we use the results of Refs.\cite{[12],[19]} for $^{131}Xe$ and $^{73}Ge$ respectively. \bigskip \begin{table} \centering \caption{ Characteristics of some current experiments. In the second column is reported the quenching factor $Q$, in the third column the electron--equivalent energy at threshold, in the fourth the square of the form factor at threshold, and in the last column the present experimental sensitivity. } \begin{tabular}{|r|r|r|r|r|} \hline {\em \mb{Nucleus}} & $Q$ & $E_{ee}^{th}(\mb{KeV})$ & $F^{2}(E_{R}^{th})$ & ${\mb{evts/(Kg~d~ KeV)}} $\\ \hline Ge\cite{[14]} & 0.25 & 2 & 0.87 & 3.0 \\ \hline Ge\cite{klapdor} & 0.25 & 12 & 0.41 & 0.2 \\ \hline Xe\cite{[17]} & 0.80 & 40 & 0.07 & 0.8 \\ \hline \end{tabular} \end{table} \subsection{Local Neutralino Density} We denote the local halo density by $\rho_l$, for which we use the estimate $\rho_l = 0.5$ GeV cm$^{-3}$ \cite{turner}. For the value of the local neutralino density $\rho_\chi$ to be used in the rate of Eq.(\ref{eq:dir2}), for each point of the model parameter space we take into account the relevant value of the cosmological neutralino relic density. When $\Omega_\chi h^2$ is larger than a minimal $(\Omega h^2)_{\rm min}$ required by observational data and by large scale structure calculations we simply put $\rho_\chi=\rho_l$. When $\Omega_\chi h^2$ turns out less than $(\Omega h^2)_{\rm min}$, the neutralino may only provide a fractional contribution ${\Omega_\chi h^2 / (\Omega h^2)_{\rm min}}\equiv \xi$ to $\Omega h^2$; in this case we take $\rho_\chi = \rho_l \xi$. The value to be assigned to $(\Omega h^2)_{\rm min}$ is somewhat arbitrary. Here we set it equal to 0.1. It is worth remarking here that, due to this scaling procedure, for the direct detection rate one has: i) $R_{0,i} \propto \rho_l \sigma_i$ for ${\Omega_\chi h^2 \geq (\Omega h^2)_{\rm min}}$ and ii) $R_{0,i} \propto \rho_l\xi\sigma_i \propto \rho_l\sigma_i/ <\sigma_{\mbi{ann}} v>_{\mbi{int}}$ for ${\Omega_\chi h^2 < (\Omega h^2)_{\rm min}}$. Thus the rate $R_{0,i}$ is large in the regions of the parameter space where $\sigma_i$ is large. This is trivial in case i), but it is also true in case ii), since when $\sigma_i$ is large also $\sigma_{\mbi{ann}}$ increases but in such a way that usually the ratio $\sigma_i/\sigma_{\mbi{ann}}$ increases too. Because of the relation $\Omega_{\chi} h^2 \propto <\sigma_{\mbi{ann}} v>^{-1}_{\mbi{int}}$ it follows that $R_{0,i}$ is large for neutralino configurations with modest values of the relic abundance, and {\em vice versa}. \subsection{Results for detection rates} The most significant quantity in comparing experimental data and theoretical evaluations for direct detection is the differential rate $dR/dE_{ee} = (dR/dE_R)/Q$ (with $dR/dE_R$ defined in Eq.(\ref{eq:dir1})) rather than the total rates, obtained by integration over wide ranges of $E_{ee}$. By using the differential rate instead of the integrated ones, one obtains the best signal--to--background ratio. Note that the experimental spectra, apart from an energy interval around threshold, usually show a very flat behaviour, whereas signals for light neutralinos are decreasing functions of the nuclear recoil energy. A complete procedure would then be to compare the experimental and theoretical rates over the whole $E_{ee}$ range. However, to simplify the presentation here, we give our results in terms of the rate integrated over a narrow range of 1 KeV at a specific value of $E_{ee}$, the one which appears the most appropriate for each experiment: typically it corresponds to a point close to the experimental threshold. To be definite we consider the following cases: i) Ge (natural composition). Among the various running experiments \cite{mosca}, we select the {two} which, at present, appear to provide the most stringent limits: a) Caltech--PSI--Neuchatel \cite{[14]} with $E^{th}_{ee} = 2$ KeV, differential rate $\simeq 3$ events/(Kg day KeV); b) Heidelberg--Moscow \cite{klapdor} with $E^{th}_{ee} = 12$ KeV, differential rate $ \simeq 0.2$ events/(Kg day KeV). Correspondingly, for Ge we have evaluated our rate by integrating $dR/dE_{ee}$ over the range (2--3) KeV for experiment a) and over (12--13) KeV for experiment b). It turns out that the case b) provides the most stringent bound also for light neutralinos. ii)$^{129}$Xe. In this case, taking into account the features of the DAMA experiment \cite{[17]}, we have considered the rate R integrated over the range (40--41) KeV. Our results are shown in Figs.16--19. Figs.16--18 report the rate for a Ge detector for the regions of the parameter space which are depicted in Figs.12--14, respectively. In parts (a) and (b) of each figure, R is displayed in the form of a scatter plot, in terms of $m_{\chi}$ and of the relic abundance, respectively. The horizontal line denotes the present level of sensitivity in the Heidelberg--Moscow experiment. We notice that, in all cases shown in these figures, the experimental sensitivity is already, for some configurations, at the level of the predicted rate. Some points of the supersymmetric parameter space, denoted by filled squares in Figs.12--14, are even already excluded by present data. The exploration potential of this class of experiments as the sensitivity is improved is apparent from these figures. Fig.19 shows the rate R for $^{129}$Xe for the region of the parameter space displayed in Fig.13: again the horizontal line gives the present experimental sensitivity. A comparison of Fig.19 with Fig.17 shows that the Ge experiments are currently more effective. However, it has to be noticed that experiments with liquid Xe may become extremely competitive in the future \cite{[17]}. A few more remarks are in order here: i) The cases displayed in Figs.16--18 present the common feature of providing fair chances for direct detection. This is not a surprise, since these representative points all belong to the category of configurations with small values of $M_A$. As was stressed before, once we move away from these appealing physical regions of the neutralino parameter space, the rates for direct detection may fall far below (by many orders of magnitude) the detection sensitivities (present or future). This unfortunate situation occurs, for instance, typically as we move towards smaller values of $\tb$. However, one should keep in mind that the regime of very large $\tb$, where signals may be sizeable, represents a very interesting scenario, deserving much attention and exploration. In fact this is one of the two options, very small or very large $\tb$, which seem to fit low--energy phenomenology at the best \cite{ch}. ii) The scatter plots in parts b) of Figs.16--19 show explicitly a property previously mentioned in Sect.IX.C, namely that the scaling procedure adopted to evaluate the neutralino local density implies a R--$\Omega_{\chi} h^2$ correlation. Configurations which provide a measurable R usually entail a low $\Omega$ and viceversa. Only in a few cases the neutralino may be detectable by direct detection and also provide a sizeable contribution to $\Omega$. \section{Conclusions} In the present paper we have discussed some possible scenarios for neutralino dark matter which originate from the relaxation of the assumption of strict universality for soft scalar masses at $M_{GUT}$. This approach derives from the general consideration that many crucial theoretical points entering not only grand unified and supersymmetric theories, not to mention the Standard Model, are far from being understood and/or verified. For this reason, any new theoretical assumption has to be fully scrutinized. This is even more important because new assumptions in supersymmetric models are often introduced not because of solid arguments, but rather for the sake of simplicity and for the need to reduce the large number of free parameters that would otherwise prevent any firm prediction. In our work we have discussed different scenarios, by considering various physical constraints in a sort of hierarchical order, giving top priority to the requirement of radiative EWSB, implemented with a no--fine--tuning criterion, and to the cosmological relic neutralino density constraint. Some other assumptions, often introduced in the literature, have been relaxed in our work. This is in particular the case for universality in the soft scalar masses. However, it has to be remarked that the type of departure from universality that we have considered in our paper is far from being the most general one, as was noticed in Sect.II. In particular, it only refers to the Higgs masses, and not to the sfermion masses. The implications of the various scenarios on neutralino relic abundances and rates for detection rates have been analysed, and the impact of a non--universality in $m_0$ has been discussed for the whole range of $\tb$. We have shown that the departure from $m_0$ universality is particularly interesting in two respects: i) Small values of $M_A$ are allowed: this has in itself the dramatic consequence for direct detection of generating a large value for the angle $\alpha$ and large couplings to matter of the lightest neutralino $\chi$. ii) Higgsino or mixed higgsino--gaugino configurations appear for all $\tb$: this contrasts with the pure gaugino configurations favoured by strict $m_0$ universality. \hfill \break Consequences of such a departure from universality on the size of the neutralino relic abundance have been analysed for both large and small values of $\tan \beta$. It has been shown that, because of the previous properties, deviations from universality may reduce the value of $\Omega_{\chi} h^2$. The predicted rates for direct detection has been analysed in detail and compared with current and foreseen experimental sensitivities. The role of the previous properties in opening interesting perspectives for this kind of search has been elucidated. We find that presently--running experiments are already impacting interesting regions of the neutralino parameters space in some of the non--universal scenarios discussed here. {\bf Acknowledgements.} We wish to express our thanks to Uri Sarid for interesting discussions and for contributions in the very early stages of our work. We also gratefully acknowledge very useful conversations with Marek Olechowski and informative discussions on experimental aspects of direct detection with Pierluigi Belli, Rita Bernabei and Luigi Mosca. NF wishes to express his gratitude to the Fondazione A. Della Riccia for a fellowship. This work was supported in part by the Ministero della Ricerca Scientifica e Tecnologica (Italy).
\section{Introduction} \indent Over the decades, as is well known, a wide variety of regularization schemes have been developed in quantum field theory \cite{1}. However, as every schemes heve their own distinct advantages and disadvantages, this topic is still one of the important and fundamental issues under investigation. One of the most challenging problem is perhaps how to preserve all properties of the original action manifestly and consistently. A few years ago, a new regularization method named intrinsic vertex regularization was first proposed for the $\phi^4$ theory by Wang and Guo \cite{2}. The key point of the method is, in fact, based upon the following simple observation: For a given ultraviolet divergent function at certain loop order in a renormalizable QFT, there always exists a set of convergent functions at the same loop order such that their Feynman graphs share the same loop skeleton and the main difference is that the convergent ones have additional vertices of certain kind and the original one is the case without these vertices. This is, in fact, a certain intrinsic relation between the original ultraviolet divergent graph and the convergent ones in the QFT. It is this relation that indicates it is possible to introduce the regularized function for the divergent function with the help of those convergent ones so that the potentially divergent integral of the graph can be rendered finite while for the limiting case of the number of the additional vertices $q\to 0$ the divergence again becomes manifest in pole(s) of $q$. To be concrete, let us consider a 1PI graph with $I$ internal lines at one loop order in the $\phi^4$ theory. Its superficial degree of divergences in the momentum space is $$ \delta=4-2I. $$ When $I=1$ or $2$, the graph is divergent. Obviously, there exists such kind of graphs that they have additional $q$ four-$\phi$-vertices in the internal lines. Then the number of internal lines in these graphs is $I+q$ so that the divergent degree of the new 1PI graphs become $$ \delta'=4-2(I+q). $$ If $q$ is large enough, the new ones are convergent and the original divergent one is the case of $q=0$. Thus, a certain intrinsic relation has been reached between the original divergent 1PI graph and the new convergent ones at the same loop order. However, application of this method to QED runs into a difficulty. The problem is that, unlike the the $\phi^4$ theory, the electron-photon vertex in QED carries a $\gamma$-matrix and is a Lorentz vector. As a result, simply inserting the vertex would increase the rank of the function as Lorentz tensors and would make the problem quite complicated. In order to overcome this difficulty, in \cite{3,4} the authors introduced an alternative method. We follow the example of the $\phi^4$ theory to demonstrate the key point of this method: One shifts the mass term of the $\phi$ field from $m^2$ to $m^2+\mu^2$ and regards $\mu^2\phi^2$ as a new vertex in addition to the vertex $\lambda\phi^4$. By inserting the new vertex into the internal $\phi$ lines in the graph of a given 1PI $n$-point divergent function, a set of new convergent functions can be obtained provided the number of inserted vertices, $q$, is large enough. Then one can introduce a new convergent function, the regularized function, and the potential infinity in the original 1PI $n$-point function may be recovered as the $q\to 0$ limiting case of that function. Obviously, the mass shifting method can be easily generalized to QED by simply shifting the electron mass from $m$ to $m+\mu$ and regarding the term $-\mu\bar{\psi}{\psi}$ as a new vertex. In fact, as has been shown in \cite{4}, it turns out to be successful to QED. Nevertheless, it is not really intrinsic since the Feynman rules of the theory have to be modified. As a result, it may not completely work for non-Abelian gauge theories, {\it e.g.}, QCD, because generally it is not clear whether the gauge symmetry can be preserved manifestly for these theories, although such a proof for QED at one loop level has been given \cite{6}. Very recently, we presented an improved approach in \cite{7} to reexamine the $\phi^4$ theory and QED, in which a new concept, {\it inserter}, was introduced. An inserter is a vertex or a pair of vertices linked by an internal line, in which the momenta of the external legs are all set to zero, and, if there are any, all the Lorentz indices are contracted in pair by the spacetime metric and all the internal gauge symmetry indices are contracted by the Killing-Cartan metric in the corresponding representation, so that as a whole an inserter always carries the vacuum quantum numbers, i.e. zero momentum, scalar in the spacetime symmetry, and singlet in internal and gauge symmetries. It is not hard to see that in any given QFT as long as a suitable kind of inserters are constructed with the help of the Feynman rules of the theory, some intrinsic relations between the divergent functions and convergent ones at the same loop order will be found by simply regarding the convergent ones as the ones given by suitably inserting $q$-inserters in all internal lines in the given divergent ones. The crucial point of this approach, therefore, is very simple but fundamental, that is, the entire procedure is intrinsic in the QFT. There is nothing changed, the action, the Feynman rules, the spacetime dimensions etc. are all the same as that in the given QFT. This is a very important property which should shed light on the challenging problem mentioned at the beginning of the paper. Consequently, in applying to other cases all symmetries and topological properties there should be preserved in principle. In what follows, we concentrate on how to apply the inserter approach to QCD at one loop order. We present the main steps and the results of the inserter regularization procedure for it. We find that, as is expected, the gauge invariance is preserved manifestly, and all results are the same as those derived by means of other regularization methods. \section{Intrinsic Regularization in QCD} \indent The QCD Lagrangian, including ghost fields and gauge fixing terms, can be written as \begin{eqnarray} {\cal L}_{QCD} & = & -\frac{1}{4} (\partial_{\mu} A^a_{\nu}-\partial_{\nu}A^a_{\mu}+g_cf^{abc}A^b_{\mu}A^c_{\nu}) (\partial^{\mu} A^{a\nu} -\partial^{\nu} A^{a\mu}+g_cf^{ade}A^{d\mu}A^{e\nu}) \nonumber \\ & & -\frac{1}{2\xi}(\partial_{\mu} A^{a\mu})^2 -\bar{\eta}^a \partial_{\mu}(\partial^{\mu} \delta^{ac}-g_cf^{abc}A^{b\mu})\eta^c +\bar{\psi}[i\gamma_{\mu}(\partial^{\mu} - ig_cA^{a\mu} \frac{\lambda^a}{2})-m]\psi~, \label{L} \end{eqnarray} and the Feynman rules are well known. The main steps of the inserter approach for QCD may be stated more concretely as follows. First, we should construct the inserters in QCD. This work, with regards to simplicity and consistency with other theories, {\it e.g.}, the electroweak theory, may actually be done within a more general framework, namely, within the framework of the standard model in which QCD is contained. The explicit expressions of all inserters in the standard model have been preestablished in \cite{7}. Here we merely list those relevant to QCD: \begin{itemize} \item The gluon-inserter: \begin{eqnarray} I^{\{g\}ab}_{~~\mu\nu}(p)=-6ig_c^2C_2({\bf 8}) g_{\mu\nu} \delta^{ab}. \label{gluon-ins} \end{eqnarray} \item The ghost-inserter: \begin{eqnarray} I^{\{gh\}}_{~~~a_1a_2}(p) = -ig_c^2 C_2({\bf 8}) \delta_{a_1a_2}. \label{ghost-ins} \end{eqnarray} \item The quark-inserter: \begin{eqnarray} I^{\{q\}}(p)=-i\lambda_q. \label{quark-ins} \end{eqnarray} \end{itemize} In eqs.(\ref{gluon-ins}) and (\ref{ghost-ins}), $C_2({\bf 8})$ is the second Casimir operator valued in the adjoint representation of $SU_c(3)$ algebra. In eq.(\ref{quark-ins}), $\lambda_q$ takes value $\frac{g}{2}\frac{m_q}{M_W}$ in the standard model, but here its value is irrelevant for our purpose. It should be mentioned that here the quark inserters are constructed by borrowing the fermion-Higgs-vertex of Yukawa type from the standard model, this is in analogy with as occurs in QED. The issue has been discussed in detail in \cite{7}. For a given divergent 1PI amplitude $\Gamma^{(n_f,n_g)}(p_1,\cdots, p_{n_f};k_1,\cdots, k_{n_g})$ at the one loop order with $n_f$ external fermion lines and $n_g$ external photon lines, we consider a set of 1PI amplitudes $\Gamma^{(n_f,n_g)} (p_1,\cdots, p_{n_f};k_1,\cdots, k_{n_g}; q)$ which correspond to the graphs with, if the loop contained in the graph purely consists of fermion lines, all possible $2q$ insertions of the fermion inserter in the internal fermion lines, or in other cases, all possible $q$ insertions of the corresponding inserter in the internal boson (ghost) lines in the original graph. The divergent degree therefore becomes: $$ \delta=4-I_f-2I_g-2q. $$ If $q$ is large enough, $\Gamma^{(n_f,n_g)}(p_1,\cdots,p_{n_f};k_1,\cdots, k_{n_g}; q)$ are convergent and the original divergent function is the case of $q=0$. Thus we reach a relation between the given divergent 1PI function and a set of convergent 1PI functions at the one loop order. In fact, the function of inserting the inserter(s) into internal lines is simply to raise the power of the propagator of the lines and to decrease the degree of divergence of given graph. In order to regularize the given divergent function with the help of this relation, we need to deal with those convergent functions on an equal footing and pay attention to their differences due to the insertions. To this end, we introduce a new function: \begin{eqnarray} \begin{array}{ll} \Gamma^{(n_f,n_g)}(p_1,\cdots, p_{n_f};k_1,\cdots, k_{n_g}; ~q;~\mu) {}~~~~~~~\\[4mm] {}~~~~~~~= (-i\mu)^{2q} (-i\lambda)^{-2q} \frac 1 {N_q} \sum \Gamma^{(n_f,n_g)}(p_1,\cdots, p_{n_f};k_1,\cdots, k_{n_g}; ~q) \end{array} \end{eqnarray} where $\mu$ is an arbitrary reference mass parameter, the summation is taken over the entire set of such ${N_q}$ inserted functions, and the factor $(-i\lambda)^{-2q}$ introduced here, in which $\lambda$ stands for $\lambda_q$ for fermion loop and for $g_c$ for other cases, is to cancel the ones coming from the inserters. It is clear that this function is the arithmetical average of those convergent functions and has the same dimension in mass, the same order in coupling constant with the original divergent 1PI function. Then we evaluate it and analytically continue $q$ from the integer to the complex number. Finally, the original 1PI function is recovered as its $q\to 0$ limiting case: \begin{eqnarray} \Gamma^{(n_f,n_g)}(p_1,\cdots, p_{n_f};k_1,\cdots, k_{n_g}) =\lim_{q\to 0} \Gamma^{(n_f,n_g)}(p_1,\cdots, p_{n_f};k_1,\cdots, k_{n_g};q;\mu), \end{eqnarray} and the original infinity appears as pole in $q$. Similarly, this procedure should work for the cases at the higher loop orders in principle. The divergent 1PI graphs at the one loop order in QCD are as follows: the gluon self-energy $\Pi_{\mu \nu}^{ab}(k)$, the quark self-energy $\Sigma(p)$, the ghost self-energy $\tilde{\Pi}^{ab}(p)$, renormalized by $Z_3$, $Z^F_3$ and $\tilde{Z}_3$, the three-gluon vertex $\Gamma^{abc}_{\mu\nu\lambda} (k_1,~k_2)$, the four-gluon vertex $\Gamma^{abcd}_{\mu\nu\lambda\tau}(k_1,~k_2,~k_3)$, the quark-gluon vertex $\Gamma^{a}_{\mu}(p^{\prime},p)$, and the ghost-gluon vertex $\tilde{\Gamma}^{abc}_{\mu}(p^{\prime},p)$, with the renormalization constant $Z_1$, $Z_4$, $Z^F_1$, $\tilde{Z}_1$. In addition, there is a mass shift for the quark, which we shall ignore. All corresponding graphs are listed in figures 1-7. As numerious diagrams are concerned, evaluating them one by one in detail would be much lengthy and unnecessary. In the next section, we will evaluate in detail the gluon self-energy $\Pi_{\mu \nu}^{ab}(k)$ as a typical example to show the main step of the approach, paying special attentions to the gauge inva riance of the function. Then in the subsequent section, we will directly give all the results corresponding to other involved diagrams to verify the Slavnov-Taylor identities at one loop order. \section{Regularization and Evaluation of the Gluon Self-Energy $\Pi_{\mu \nu}^{ab}(k)$} \indent Before the detailed evaluations are presented, we should first refer to a special problem which arises in any massless theories, {\it i.e.}, the genuine infrared divergence in these theories. In the regularization schemes, this problem usually appears as the lack of consistent definitions of the regularized Feynman integrals for the ones which are both ultraviolet and infrared divergent. For instance, in the dimensional regularization scheme, let's consider the massive integral \begin{eqnarray} \int\frac{d^{2\omega}l}{(2\pi)^{2\omega}}\frac{1}{(l^2+m^2)^n} =\frac{i\Gamma (n-\omega)}{(4\pi)^{\omega}\Gamma (n)(m^2)^{n-\omega}} \equiv I(m,~\omega,~n)~, ~~~(~m^2\neq 0~) \label{massive} \end{eqnarray} which converges for $\omega$ complex; the parameter $n$ is arbitrary but fixed. We note first of all that the limit $\displaystyle\lim_{m^2\to0}I(m,~\omega,~n)$ may or may not exist, depending on the relative magnitudes of $n$ and $\omega$. But even if it did exist, another problem could arise as we approach four-space (provided the original integral is infrared divergent to begin with), because in general $$ \displaystyle\lim_{\omega\to2} [ \displaystyle\lim_{m^2\to0} I(m,~\omega,~n)] \neq \displaystyle\lim_{m^2\to0} [\displaystyle\lim_{\omega\to2} I(m,~\omega,~n)], $$ so that the massless integral $\displaystyle\lim_{\omega \to 2}{\int\frac{d^{2\omega}l}{(2\pi)^{2\omega}}\frac{1}{(l^2)^n}}$ can not be derived unambiguously from the massive integral (\ref{massive}). Furthermore, the trick of inserting a finite mass into the integral and then allowing it to approach to zero at the end of the calculation is, in general, not a satisfactory prescription yet, because it spoils the gauge symmetry in the original theory, provided such a symmetry existed in the first place. To avoid this difficulty, 't Hooft and Veltman naively comjectured that \begin{eqnarray} \displaystyle\lim_{\omega \to 2}{\int\frac{d^{2\omega}l}{(2\pi)^{2\omega}} \frac{1}{(l^2)^n}}=0~,~~~for~ \omega,~n~complex. \label{massless} \end{eqnarray} It has been shown that no inconsistencies occur, {\it e.g.}, in the Slavnov-Taylor identities \cite{10}, due to the acceptance of the above conjecture. In our application of the present approach to QCD, as we will see, the same problem occurs, {\it e.g.}, in calculating the tadpole diagram Fig.1d of the gluon self-energy. To solve this problem, we employ a conjecture analogous to 't Hooft and Veltman's: \begin{eqnarray} \displaystyle\lim_{q\to 0}{\int\frac{d^4l}{(2\pi)^4} \frac{1}{[(k-l)^2]^{Aq}(l^2)^{Bq+n}}}=0,~~for~q,~n,~complex,~A\geq 0,~B\geq 0,~A+B=1~. \label{conj} \end{eqnarray} Likewise, we will see that no inconsistencies occur due to the acceptance of the eq.(\ref{conj}). Now we turn to evaluate in detail the gluon self-energy $\Pi_{\mu\nu}^{ab}(k)$ shown in Fig.1. The diagrams contributing to $\Pi_{\mu\nu}^{ab}(k)$ are four in number, namely, the gluon loop diagram, the ghost loop diagram, the quark loop diagrams, and the gluon tadpole. The integral expressions of the regularized diagrams in the momentum space are given in the appendix ( For simplicity, we take the Feynman gauge $\xi =1$. ). First, we consider the gluon loop contribution. From (\ref{glse-a}), a little bit of algebra yields \begin{eqnarray} \Pi_{(A)\mu\nu}^{~~~ab}(k;q;\mu)~=~-\frac{1}{2}g_c^{2}[-6C_2({\bf 8})\mu^{2}]^q \delta^{ab}I_1 ~, \nonumber \end{eqnarray} with \begin{eqnarray} I_1~=~\frac{1}{N_q} \displaystyle{\sum_{i=0}^{q} \int\frac{d^4p}{(2\pi)^4}} \frac{-2p^2g_{\mu\nu}-(5k^2-2p\cdot k)g_{\mu\nu} -10p_{\mu}p_{\nu}+2k_{\mu}k_{\nu}+5p_{\mu}k_{\nu}+5k_{\mu}p_{\nu}}{ (p^2)^{i+1}[(p-k)^2]^{q-i+1}}~. \nonumber \end{eqnarray} In the present case, $N_q=q+1$. Note that because of eq.(\ref{conj}), the contribution of the first term in the numerator of the above equation actually vanishes, so it can be neglected. Using the Feynman parameterization, we get \begin{eqnarray} I_1 & = & \frac{1}{q+1} \displaystyle{\sum_{i=0}^{q}} \frac{\Gamma (q+2)}{\Gamma (i+1)\Gamma (q-i+1)} \displaystyle{\int_0^1dx} x^{q-i}(1-x)^i \nonumber \\ & & \times \displaystyle{\int\frac{d^4p}{(2\pi)^4}} \frac{(2x-5)k^2g_{\mu\nu} -10p_{\mu}p_{\nu} -(10x^2-10x-2)k_{\mu}k_{\nu}}{[p^2+x(1-x)k^2]^{q+2}}~, \label{I1}\end{eqnarray} where we have made a momentun shift: $p \rightarrow p-kx$, and the linear terms in $p$ in the numerator have been dropped since they do not contribute to the integral. Now the integration over $p$ can be performed by using the following formulas: \renewcommand{\theequation}{\arabic{equation}\alph{abc}} \setcounter{abc}{0} \addtocounter{abc}{1} \begin{eqnarray} \int \frac{d^4 p}{(2\pi)^4} \frac{(p^2)^{\beta}}{(p^2+M^2)^A} & = & \frac{i}{(4\pi)^2} \frac{\Gamma(2+\beta)\Gamma(A-2-\beta)}{\Gamma(A)}(M^2)^{2+\beta-A}, \label{int-a} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \vspace{-6mm} \begin{eqnarray} \int \frac{d^4 p}{(2\pi)^4} \frac{(p^2)^{\beta}p_{\mu}p_{\nu}}{(p^2+M^2)^A} & = & \frac{i}{(4\pi)^2}\frac{1}{4}g_{\mu\nu} \frac{\Gamma(3+\beta)\Gamma(A-3-\beta)}{\Gamma(A)}(M^2)^{3+\beta-A}, \label{int-b} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \vspace{-6mm} \begin{eqnarray} \int \frac{d^4 p}{(2\pi)^4} \frac{(p^2)^{\beta}p_{\mu}p_{\nu}p_{\rho}p_{\sigma}} {(p^2+M^2)^A} & = & \frac{i}{(4\pi)^2}\frac{1}{24}(g_{\mu\nu}g_{\rho\sigma}+g_{\mu\rho}g_{\nu\sigma} +g_{\mu\sigma}g_{\nu\rho}) \nonumber \\ & & \times \frac{\Gamma(4+\beta)\Gamma(A-4-\beta)}{\Gamma(A)}(M^2)^{4+\beta-A}. \label{int-c} \end{eqnarray} \addtocounter{abc}{1} \renewcommand{\theequation}{\arabic{equation}} {}From (\ref{I1}), (\ref{int-a}), and (\ref{int-b}) we get: $$ \begin{array}{ll} I_1~ = & \frac{i}{(4\pi)^2} \displaystyle{\sum_{i=0}^{q}} \frac{\Gamma (q-1)}{(q+1)\Gamma (i+1)\Gamma (q-i+1)} \displaystyle{\int_0^1dx} x^{q-i}(1-x)^i \nonumber \\ & \times \displaystyle{ \frac{(q-1)[(2x-5)k^2g_{\mu\nu}-(10x^2-10x-2)k_{\mu}k_{\nu}] -5x(1-x)k^2g_{\mu\nu}}{[x(1-x)k^2]^{q}} } \nonumber \\ ~~~~~ = & \frac{i}{(4\pi)^2} \frac{\Gamma (q-1)}{\Gamma (q+2)} \displaystyle{\int_0^1dx} \frac{[q(2x-5)+(5x^2-7x+5)]k^2g_{\mu\nu}-(q-1)(10x^2-10x-2)k_{\mu}k_{\nu}} {[x(1-x)k^2]^{q}}~, \nonumber \end{array} $$ where in the last step the summation over $i$ has been performed with the help of the binomial theorem. This expression also makes sense when we make an analytical continuation of $q$ from integer to complex number. When $q\to 0$, using the expansion $$ \frac{\Gamma(q-1)}{\Gamma(q+2)}=-\frac{1}{q}-q+O(q^2),~~~ [x(1-x)k^2]^q=1+q\ln [x(1-x)k^2]+O(q^2)~, $$ and making a rescaling of the parameter $\mu^2 \rightarrow -6C_2({\bf 8})\mu^2$, we get: \begin{eqnarray} \Pi_{(A)\mu\nu}^{~~~ab}(k;q;\mu) & = & \frac{ig_c^2C_2({\bf 8})}{(4\pi)^2} \delta^{ab} [ (\frac{19}{12}k^2g_{\mu\nu}-\frac{11}{6}k_{\mu}k_{\nu})\frac{1}{q} \nonumber \\ & & -(\frac{19}{12}k^2g_{\mu\nu}-\frac{11}{6}k_{\mu}k_{\nu})\ln\Bigl(\frac{k^2}{\mu^2}\Bigr)+(\frac{47}{36}k^2g_{\mu\nu}-\frac{14}{9}k_{\mu}k_{\nu}) ]~. \label{no13} \end{eqnarray} Clearly $\Pi_{(A)\mu\nu}^{~~~ab}(k;q;\mu)$ does not conserve current, this is due to our choice to use a covariant gauge (rather than, for example, an axial gauge). For the sake of this choice, we have to introduce spurious gluon polarization states. These spurious states must be removed by taking the ghost loop contribution into account. We could have computed with an axial or ``ghost free'' gauge, but it is usually much easier to use the simple Feynman gauge and add in the ghost contribution. To compute ghost loop contribution (\ref{glse-b}), we use the same Feynman parameterization arriving at \begin{eqnarray} \Pi_{(B)\mu\nu}^{~~~ab}(k;q;\mu) ~=~ -g_c^2 C_2({\bf 8})[C_2({\bf 8})\mu^2]^q \delta^{ab}I_2 ~, \nonumber \end{eqnarray} with \begin{eqnarray} I_2 ~=~ \displaystyle{{\int_0^1dx} \int\frac{d^4p}{(2\pi)^4}} \frac{p_{\mu}p_{\nu}-x(1-x)k_{\mu}k_{\nu}}{[p^2+x(1-x)k^2]^{q+2}}~, \nonumber \end{eqnarray} where momentun shift: $p \rightarrow p-kx$ has been made, and the linear terms in $p$ in the numerator have been dropped. After performing the integration over $p$ and taking the limit $q\to 0$ subsequently, we obtain \begin{eqnarray} \Pi_{(B)\mu\nu}^{~~~ab}(k;q;\mu)& = & \frac{ig_c^2C_2({\bf 8})}{(4\pi)^2} \delta^{ab} [ (\frac{1}{12}k^2g_{\mu\nu}+\frac{1}{6}k_{\mu}k_{\nu})\frac{1}{q} \nonumber \\ & & -(\frac{1}{12}k^2g_{\mu\nu}-\frac{1}{6}k_{\mu}k_{\nu})\ln\Bigl(\frac{k^2}{\mu^2}\Bigr)+\frac{5}{36}k^2g_{\mu\nu}+\frac{1}{9}k_{\mu}k_{\nu}) ]~. \label{no14} \end{eqnarray} Again, we find that $\Pi_{(B)\mu\nu}^{~~~ab}(k;q;\mu)$ also does not conserve current. However, the non-conserving term in (\ref{no14}) exactly cancels that in (\ref{no13}), and makes $\Pi_{(A)\mu\nu}^{~~~ab}(k;q;\mu)+\Pi_{(B)\mu\nu}^{~~~ab}(k;q;\mu)$ gauge invariant, namely, $$ k^{\mu}[\Pi_{(A)\mu\nu}^{~~~ab}(k;q;\mu)+\Pi_{(B)\mu\nu}^{~~~ab}(k;q;\mu)] =0~. $$ Next, the quark loop contribution $\Pi_{(C)\mu\nu}^{~~~ab}(k;q;\mu)$ need not be recalculated since it is just $\Pi_{\mu\nu}^{QED}(k;q;\mu)$, which can be found in \cite{6,7}, multiplied by a color factor $N_fTr({\bf T}^a{\bf T}^b)$: \begin{eqnarray} \Pi_{(C)\mu\nu}^{~~~ab}(k;q;\mu) & = & N_fTr({\bf T}^a{\bf T}^b) \Pi_{\mu\nu}^{QED}(k;q;\mu)=\frac{4ig_c^2N_fC_2({\bf 3})}{(4\pi)^2} \delta^{ab} (k_{\mu}k_{\nu}-k^2g_{\mu\nu}) \nonumber \\ & & \times [\frac{1}{3q}+\frac{2}{3}-2\int_{0}^{1} dx x(1-x)\ln\frac{k^2 x(1-x)-m^2} {\mu^2}+\cdots]~. \label{no15} \end{eqnarray} Obviously, $\Pi_{(C)\mu\nu}^{~~~ab}(k;q;\mu)$ itself is gauge invariant. Finally, as in dimensional regularization scheme, from eq.(\ref{conj}) we know that the contribution of the gluon tadpole diagram vanishes: \begin{eqnarray} \Pi_{(D)\mu\nu}^{~~~ab}(k;q;\mu)=0~. \label{no16} \end{eqnarray} {}From (\ref{no13}), (\ref{no14}), (\ref{no15}), and (\ref{no16}) we obtain the final expression of the regularized gluon self-energy $\Pi_{\mu\nu}^{ab}(k;q;\mu)$: \begin{eqnarray} \Pi_{\mu\nu}^{ab}(k;q;\mu) & = & \Pi_{(A)\mu\nu}^{~~~ab}(k;q;\mu) +\Pi_{(B)\mu\nu}^{~~~ab}(k;q;\mu)+\Pi_{(C)\mu\nu}^{~~~ab}(k;q;\mu) \nonumber \\ & = & \frac{ig_c^2}{(4\pi)^2}\delta^{ab}(k^2g_{\mu\nu}-k_{\mu}k_{\nu}) \{[\frac{5}{3}C_2({\bf 8})-\frac{4}{3}N_fC_2({\bf 3})]\frac{1}{q}+\cdots\}~, \end{eqnarray} which is the same as that derived by means of other regularization methods and satisfies the gauge invariance condition $k^{\mu}\Pi_{\mu\nu}^{ab}(k;q;\mu)=0$. \section{Renormalization Constants and Slavnov-Taylor Identities at one Loop Order} \indent So far we have calculated the gluon self-energy $\Pi_{\mu\nu}^{ab}(k)$ in detail by means of the intrinsic regularization method at one loop level. As we have expected, the result turns out to be gauge invariant. However, this is not enough for us to say that the method preserves the gauge invariance of QCD. As a complement, we should further show that the Slavnov-Taylor identities between the renormalization constants hold, namely, \begin{eqnarray} \frac{Z_1}{Z_3} ~=~ \frac{Z^F_1}{Z^F_3} ~=~ \frac{\tilde{Z}_1}{\tilde{Z}_3} ~=~ \frac{Z_4}{Z_1}~. \label{stid} \end{eqnarray} It is these identities that guarantee that the renormalized theory possess the same gauge theory structure as the original one. Moreover, they are essential for proving the renormalizability and unitarity of the theory. To verify these identities, all divergent 1PI graphs, not only $\Pi_{\mu\nu}^{ab}(k)$, but also others, must be taken into account. After lengthy and tedious calculations, we get \renewcommand{\theequation}{\arabic{equation}\alph{abc}} \setcounter{abc}{0} \addtocounter{abc}{1} \vspace{-6mm} \begin{eqnarray} Z_1=1+\frac{g_c^2}{(4\pi)^2}[\frac{2}{3}C_2({\bf 8})-\frac{4}{3}N_fC_2({\bf 3})]\frac{1}{q} \label{const-a} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \vspace{-6mm} \begin{eqnarray} Z_3=1+\frac{g_c^2}{(4\pi)^2}[\frac{5}{3}C_2({\bf 8})-\frac{4}{3}N_fC_2({\bf 3})]\frac{1}{q} \label{const-b} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \vspace{-6mm} \begin{eqnarray} Z_4=1+\frac{g_c^2}{(4\pi)^2}[-\frac{1}{3}C_2({\bf 8})-\frac{4}{3}N_fC_2({\bf 3})]\frac{1}{q} \label{const-c} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \vspace{-6mm} \begin{eqnarray} Z^F_1=1-\frac{g_c^2}{(4\pi)^2}[C_2({\bf 8})+\frac{8}{3}C_2({\bf 3})]\frac{1}{q} \label{const-d} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \vspace{-6mm} \begin{eqnarray} Z^F_3=1-\frac{g_c^2}{(4\pi)^2}\frac{8}{3}C_2({\bf 3})\frac{1}{q} \label{const-e} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \vspace{-6mm} \begin{eqnarray} \tilde{Z}_1=1-\frac{g_c^2}{(4\pi)^2}\frac{C_2({\bf 8})}{2}\frac{1}{q} \label{const-f} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \vspace{-6mm} \begin{eqnarray} \tilde{Z}_3=1+\frac{g_c^2}{(4\pi)^2}\frac{C_2({\bf 8})}{2}\frac{1}{q} \label{const-g} \end{eqnarray} \addtocounter{abc}{1} \renewcommand{\theequation}{\arabic{equation}} It is easy to see that the Slavnov-Taylor identities (\ref{stid}) indeed holds. For a close observation, we note that although these expressions are calculated in a specific gauge therefore their gauge dependence are not explicit, they are in fact all gauge dependent ( For instance, in the axial gauge, the effective Lagrantian has no ghosts and has the same structure as in QED, resulting in the identity $Z^F_1=Z^F_3$, which is patently untrue in the Feynman gauge.). We also remark that the fermion contribution to the vector boson quartic self-interaction must diverge if the Slavnov-Taylor identities (\ref{stid}) are to hold because we explicitly see that the ratio $Z^F_1/Z^F_3Z_3^{\frac{1}{2}}$, by which the bare coupling constant is related to the renormalized one, contains a fermion contribution. On the contrary, the corresponding box diagram of QED is finite. The reason is that in QED the box diagram's divergence vanishes only upon symmetrization of the external photon lines denoted only by their vector in dices, while in QCD the symmetrization of the external lines can be performed in two ways: by symmetrizing on both vector and group indices, which as in QED gives no divergence, {\it or} by antisymmetrizing on both vector and group indices. It is this new contribution which deverges. The same reason can be applied to the fermion contribution to the triple gauge vertex. \section{Concluding Remarks} \indent We have shown the main steps and results for the regularization of the divergent 1PI functions at the one loop order in QCD by means of the inserter proposal for the intrinsic regularization method. It turns out to be satisfactory: The gauge invariance is preserved manifestly and the results are the same as those derived by means of other regularization methods. It is natural to expect that this proposal should be available to the cases at higher loop orders in principle. The renormalization of the QCD under consideration in this scheme should be the same as in usual approaches. Namely, we may subtract the divergent part of the $n$-point functions at each loop order by adding the relevant counterterms to the action. The renormalized $n$-point functions are then evaluated from the renormalized action. In the limiting case, we get the finite results for all correlation functions. It should be mentioned that the method presented here is somewhat analogous to the analytic regularization method developed by Speer \cite{11,12}. However, the two methods are in fact different: As is well known, the analytic regularization method violates unitarity, this is due to the fact that it actually continuing the power of propagators in an arbitrary way. While in our method this not the case. Here we trie to find out a procedure of regularization from some physical principles. That is, giving a ultraviolet divergent process, one can always find a set of convergent function obtainable from existing Feymann rules and for the limiting case it turns to be the original ultra-divergent one. Or from another point of view, given a ultra-divergent function, one can always extract a set of convergent functions from a certain physical process, which is gauge invariant. After taking the limitation of the convergent functions, the divergent function is naturally regularized. There is nothing changed, the action, the Feynman rules, the spacetime dimensions etc. are all the same as that in the given QFT. From this viewpoint, one should have no doubt of gauge invariance and the unitarity of the method since the sum of the convergent functions comes from a certain physical process. Application of our approach to gauge theories containing spontaneous symmetry breaking such as the standard model should be straightforward. Also, it will be much helpful to apply our approach to some other cases, such as anomalies, SUSY theories {\it etc.}, since in these cases the symmetries and topological properties are sensitive to the spacetime dimensions and the number of fermionic degrees of freedom {\it etc.}, thus we are unable to tackle them consistently by means of the hitherto well-known regularization methods such as dimensional regularization method. It is reasonable to expect that the approach presented here should be able to get rid of those problems. We will investigate these issues in detail elsewhere. \bigskip \bigskip {\raggedright{\Large\bf Appendix: Integral Expressions of the Divergent 1PI Graphs at}}\\ {\raggedright{\Large\bf \* \* \* \* \* \* \* \* \* \* \* \* \* \* \* One Loop Level in QCD}}\\ \indent There are a number of divergent iPI graphs at one loop level in QCD, which contribute to the gluon self-energy $\Pi_{\mu \nu}^{ab}(k)$, the quark self-energy $\Sigma(p)$, the ghost self-energy $\tilde{\Pi}^{ab}(p)$, the three-gluon vertex $\Gamma^{abc}_{\mu\nu\lambda} (k_1,~k_2)$, the four-gluon vertex $\Gamma^{abcd}_{\mu\nu\lambda\tau}(k_1,~k_2,~k_3)$, the quark-gluon vertex $\Gamma^{a}_{\mu}(p^{\prime},p)$, and the ghost-gluon vertex $\tilde{\Gamma}^{abc}_{\mu}(p^{\prime},p)$ respectively. In what follows we present the integral expressions of the regularized diagrams in the momentum space (in Feynman gauge $\xi =1$). \noindent 1. \hspace{0.3cm}The integral expressions of the regularized diagrams contributing to $\Pi_{\mu \nu}^{ab}(k)$: \renewcommand{\theequation}{\arabic{equation}\alph{abc}} \setcounter{abc}{0} \addtocounter{abc}{1} \begin{eqnarray} \Pi_{(A)\mu\nu}^{~~~ab}(k;q;\mu)& = & \frac{1}{2}g_c^{2-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q Tr({\bf F}^a{\bf F}^b) \frac{1}{N_q}\sum_{i=0}^{q} \int\frac{d^4p}{(2\pi)^4} \nonumber \\ & & \times G_{\mu\rho_1\sigma_1}(k,-p,p-k) G_{\nu\rho_2\sigma_2}(-k,p,k-p) \nonumber \\ & & \times g^{\rho_1\rho_2} g^{\sigma_1\sigma_2} \Bigl(\frac{-i}{p^2}\Bigr)^{i+1} \Bigl(\frac{-i}{(p-k)^2}\Bigr)^{q-i+1}~, \label{glse-a} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{eqnarray} \Pi_{(B)\mu\nu}^{~~~ab}(k;q;\mu)& = & g_c^{2-2q} \mu^{2q}[-ig_c^2C_2({\bf 8})]^q Tr({\bf F}^a{\bf F}^b) \frac{1}{N_q}\sum_{i=0}^{q} \int\frac{d^4p}{(2\pi)^4} \nonumber \\ & & \times p_{\mu}(p-k)_{\nu}\Bigl(\frac{-i}{p^2}\Bigr)^{i+1} \Bigl(\frac{-i}{(p-k)^2}\Bigr)^{q-i+1}~, \label{glse-b} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{eqnarray} \Pi_{(C)\mu\nu}^{~~~ab}(k;q;\mu)& = & g_c^{2}\mu^{2q} (-i\lambda_q)^{2q} N_f Tr({\bf T}^a{\bf T}^b)\frac{1}{N_q}\sum_{i=0}^{2q} \int\frac{d^4p}{(2\pi)^4} \nonumber \\ & & \times Tr[\gamma_{\mu}\Bigl(\frac{1}{\not{p}-\not{k}-m}\Bigr)^{i+1} \gamma_{\nu}\Bigl(\frac{1}{ \not {p} -m}\Bigr)^{2q-i+1}]~, \label{glse-c} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{eqnarray} \Pi_{(D)\mu\nu}^{~~~ab}(k;q;\mu)& = & -\frac{1}{2}ig_c^{2-2q}\mu^{2q}[-6ig_c^2C_2({\bf 8})]^q [f^{abe}f^{cde} (g_{\mu\sigma}g_{\nu\rho}-g_{\mu\rho}g_{\nu\sigma}) \nonumber \\ & & +f^{ace}f^{dbe}(g_{\mu\rho}g_{\nu\sigma}-g_{\mu\nu}g_{\rho\sigma})+ f^{ade}f^{bce}(g_{\mu\nu}g_{\rho\sigma}-g_{\mu\sigma}g_{\nu\rho})] \nonumber \\ & & \times g^{\sigma\rho}\delta^{cd} \int\frac{d^4p}{(2\pi)^4} \Bigl(\frac{-i}{p^2}\Bigr)^{q+1}~, \label{glse-d} \end{eqnarray} \addtocounter{abc}{1} \renewcommand{\theequation}{\arabic{equation}} where $N_f$ is the number of flavors, ${\bf T}^a$ are generators of $SU_c(3)$ in fundamental representation, $f^{abc}$ denote the structure constants of $SU_c(3)$, $({\bf F}^a)^{bc}=-if^{abc}$ are generators of $SU_c(3)$ in adjoint representation, and $$ G_{\mu\nu\lambda}(p_1,p_2,p_3)~=~(p_1-p_2)_{\lambda}g_{\mu\nu}+(p_2-p_3)_{\mu}g_{\nu\lambda}+(p_3-p_1)_{\nu}g_{\lambda\mu} $$ comes from the three-gluon vertex. \bigskip \noindent 2. \hspace{0.3cm}The integral expressions of the regularized diagrams contributing to $\Gamma^{abc}_{\mu\nu\lambda} (k_1,~k_2)$: \renewcommand{\theequation}{\arabic{equation}\alph{abc}} \setcounter{abc}{0} \addtocounter{abc}{1} \begin{equation} \begin{array}{l} \displaystyle{\Gamma_{(A)\mu\nu\lambda}^{~~~abc}(k_1,~k_2;q;\mu) ~ = ~ ig_c^{3-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q Tr({\bf F}^a{\bf F}^b{\bf F}^c) \frac{1}{N_q}\sum_{i=0}^{q} \sum_{j=0}^{q-i} \int\frac{d^4p}{(2\pi)^4} } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{g^{\rho_1\sigma_2} g^{\rho_2\sigma_3} g^{\rho_3\sigma_1} G_{\nu\rho_2\sigma_2}(k_2,-k_2-p,p) G_{\lambda\rho_3\sigma_3}(-k_1-k_2,k_1-p,k_2+p) } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{G_{\mu\rho_1\sigma_1}(k_1,-p,p-k_1) \Bigl(\frac{-i}{p^2}\Bigr)^{i+1} \Bigl(\frac{-i}{(p+k_2)^2}\Bigr)^{j+1} \Bigl(\frac{-i}{(p-k_1)^2}\Bigr)^{q-i-j+1} }~, \end{array} \label{3g-a} \end{equation} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{equation} \begin{array}{l} \displaystyle{ \Gamma_{(B)\mu\nu\lambda}^{~~~abc}(k_1,~k_2;q;\mu) ~ = ~ -\frac{1}{2}ig_c^{3-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q \frac{1}{N_q}\sum_{i=0}^{q} \int\frac{d^4p}{(2\pi)^4}f^{as_1t_1} } \nonumber \\ {}~~~~~~~~~~~~~ \displaystyle{ G_{\mu\rho_1\sigma_1}(k_1,-p,p-k_1) [f^{bcr}f^{t_2s_2r}(g_{\nu\sigma_2}g_{\lambda\rho_2}-g_{\nu\rho_2}g_{\lambda\sigma_2}) } \nonumber \\ {}~~~~~~~~~~~~~ \displaystyle{+f^{bt_2r}f^{s_2cr}(g_{\nu\rho_2}g_{\lambda\sigma_2}-g_{\nu\lambda}g_{\rho_2\sigma_2})+f^{bs_2r}f^{ct_2r}(g_{\nu\lambda}g_{\rho_2\sigma_2}-g_{\nu\sigma_2}g_{\lambda\rho_2})] } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{\delta^{s_1s_2} \delta^{t_1t_2} g^{\rho_1\rho_2} g^{\sigma_1\sigma_2} \Bigl(\frac{-i}{p^2}\Bigr)^{i+1} \Bigl(\frac{-i}{(p-k_1)^2}\Bigr)^{q-i+1} } \nonumber \\ {}~~~~~~~~~~~~~ +\{(\mu,a,k_1) \leftrightarrow (\nu,b,k_2)\}+\{(\mu,a,k_1)\leftrightarrow (\lambda,c,k_3)\}~, \end{array} \label{3g-b} \end{equation} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{equation} \begin{array}{l} \displaystyle{ \Gamma_{(C)\mu\nu\lambda}^{~~~abc}(k_1,~k_2;q;\mu) ~ = ~ -ig_c^{3-2q} \mu^{2q}[-ig_c^2C_2({\bf 8})]^q Tr({\bf F}^a{\bf F}^b{\bf F}^c) \frac{1}{N_q}\sum_{i=0}^{q} \sum_{j=0}^{q-i} \int\frac{d^4p}{(2\pi)^4} } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{p_{\mu}(p+k_2)_{\nu}(p-k_1)_{\lambda} \Bigl(\frac{i}{p^2}\Bigr)^{i+1}\Bigl(\frac{i}{(p+k_2)^2}\Bigr)^{j+1} \Bigl(\frac{i}{(p-k_1)^2}\Bigr)^{q-i-j+1} } \nonumber \\ {}~~~~~~~~~~~~~ +\{(\nu,b,k_2) \leftrightarrow (\lambda,c,k_3)\}~, \end{array} \label{3g-c} \end{equation} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{equation} \begin{array}{l} \displaystyle{ \Gamma_{(D)\mu\nu\lambda}^{~~~abc}(k_1,~k_2;q;\mu) ~ = ~ ig_c^{3}\mu^{2q} (-i\lambda_q)^{2q} Tr({\bf T}^a{\bf T}^b{\bf T}^c) \frac{1}{N_q} \sum_{i=0}^{2q} \sum_{j=0}^{2q-i} \int\frac{d^4p}{(2\pi)^4} }~~~~~~~~~~ \nonumber\\ {}~~~~~~~~~~~~~ \times \displaystyle{Tr[ \gamma_{\mu}\Bigl(\frac{i}{\not{p}-\not{k}_1-m}\Bigr)^{i+1} \gamma_{\lambda}\Bigl(\frac{i}{\not{p}+\not{k}_2-m}\Bigr)^{j+1} \gamma_{\nu}\Bigl(\frac{i}{\not{p}-m}\Bigr)^{2q-i-j+1}] } \nonumber \\ {}~~~~~~~~~~~~~ +\{(\nu,b,k_2) \leftrightarrow (\lambda,c,k_3)\}~, \end{array} \label{3g-d} \end{equation} \addtocounter{abc}{1} \renewcommand{\theequation}{\arabic{equation}} \bigskip \noindent 3. \hspace{0.3cm}The integral expressions of the regularized diagrams contributing to $\Gamma^{abcd}_{\mu\nu\lambda\tau}(k_1,~k_2,~k_3)$: \renewcommand{\theequation}{\arabic{equation}\alph{abc}} \setcounter{abc}{0} \addtocounter{abc}{1} \begin{equation} \begin{array}{l} \displaystyle{ \Gamma_{(A)\mu\nu\lambda\tau}^{~~~~abcd}(k_1,~k_2,~k_3;q;\mu) ~ = ~ g_c^{4-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q Tr({\bf F}^a{\bf F}^b{\bf F}^c{\bf F}^d) \frac{1}{N_q}\sum_{i=0}^{q} \sum_{j=0}^{q-i} \sum_{l=0}^{q-i-j} } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{ \int\frac{d^4p}{(2\pi)^4} G_{\mu\rho_1\sigma_1}(k_1,-p,p-k_1) G_{\nu\rho_2\sigma_2}(k_2,-k_2-p,p) } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{ G_{\tau\rho_4\sigma_4}(-k_1-k_2-k_3,k_1-p,k_2+k_3+p) } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{ G_{\lambda\rho_3\sigma_3}(k_3,-k_2-k_3-p,k_2+p) g^{\rho_1\sigma_2} g^{\rho_2\sigma_3} g^{\rho_3\sigma_4} g^{\rho_4\sigma_1} } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{ \Bigl(\frac{-i}{p^2}\Bigr)^{i+1} \Bigl(\frac{-i}{(p+k_2)^2}\Bigr)^{j+1} \Bigl(\frac{-i}{(p+k_2+k_3)^2}\Bigr)^{l+1}\Bigl(\frac{-i}{(p-k_1)^2}\Bigr)^{q-i-j-l+1} } \nonumber \\ {}~~~~~~~~~~~~~ +\{(\nu,b,k_2) \rightarrow (\lambda,c,k_3),~(\lambda,c,k_3) \rightarrow (\tau,d,k_4),~(\tau,d,k_4) \rightarrow (\nu,b,k_2)\} \nonumber \\ {}~~~~~~~~~~~~~ +\{(\nu,b,k_2) \rightarrow (\tau,d,k_4)),~(\tau,d,k_4) \rightarrow (\lambda,c,k_3),~(\lambda,c,k_3) \rightarrow (\nu,b,k_2)\}~, \end{array} \label{4g-a} \end{equation} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{equation} \begin{array}{l} \displaystyle{ \Gamma_{(B)\mu\nu\lambda\tau}^{~~~~abcd}(k_1,~k_2,~k_3;q;\mu) ~ = ~ -\frac{1}{2} g_c^{4-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q \frac{1}{N_q}\sum_{i=0}^{q} \int\frac{d^4p}{(2\pi)^4} } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{ [f^{adr_1}f^{t_1s_1r_1}(g_{\mu\sigma_1}g_{\tau\rho_1}-g_{\mu\rho_1}g_{\tau\sigma_1})+f^{at_1r_1}f^{s_1dr_1}(g_{\mu\rho_1}g_{\tau\sigma_1}-g_{\mu\tau}g_{\rho_1\sigma_1}) } \nonumber \\ {}~~~~~~~~~~~~~ \displaystyle{ +f^{as_1r_1}f^{dt_1r_1}(g_{\mu\tau}g_{\rho_1\sigma_1}-g_{\mu\sigma_1}g_{\tau\rho_1})][f^{bcr_2}f^{t_2s_2r_2}(g_{\nu\sigma_2} g_{\lambda\rho_2}- g_{\nu\rho_2}g_{\lambda\sigma_2}) } \nonumber \\ {}~~~~~~~~~~~~~ \displaystyle{ +f^{bt_2r_2}f^{s_2cr_2}(g_{\nu\rho_2}g_{\lambda\sigma_2}-g_{\nu\lambda}g_{\rho_2\sigma_2})+f^{bs_2r_2}f^{ct_2r_2}(g_{\nu\lambda}g_{\rho_2\sigma_2}-g_{\nu\sigma_2}g_{\lambda\rho_2})] } \nonumber\\ {}~~~~~~~~~~~~~ \times \displaystyle{ \delta^{s_1s_2}\delta^{t_1t_2} g^{\rho_1\rho_2} g^{\sigma_1\sigma_2} \Bigl(\frac{-i}{p^2}\Bigr)^{i+1} \Bigl(\frac{-i}{(p+k_2+k_3)^2}\Bigr)^{q-i+1} } \nonumber \\ {}~~~~~~~~~~~~~ +\{(\tau,d,k_4) \leftrightarrow (\lambda,c,k_3)\}+\{(\tau,d,k_4) \leftrightarrow (\nu,b,k_2)\}~, \end{array} \label{4g-b} \end{equation} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{equation} \begin{array}{l} \displaystyle{ \Gamma_{(C)\mu\nu\lambda\tau}^{~~~~abcd}(k_1,~k_2,~k_3;q;\mu) ~ = ~ -ig_c^{4-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q \frac{1}{N_q}\sum_{i=0}^{q} \sum_{j=0}^{q-i} \int\frac{d^4p}{(2\pi)^4} } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{ \delta^{s_1t_2}\delta^{s_2t_3}\delta^{s_3t_1} [f^{adr}f^{t_1s_1r}(g_{\mu\sigma_1}g_{\tau\rho_1}-g_{\mu\rho_1}g_{\tau\sigma_1}) } \nonumber \\ {}~~~~~~~~~~~~~ \displaystyle{ +f^{at_1r}f^{s_1dr}(g_{\mu\rho_1}g_{\tau\sigma_1}-g_{\mu\tau}g_{\rho_1\sigma_1}) +f^{as_1r}f^{dt_1r}(g_{\mu\tau}g_{\rho_1\sigma_1}-g_{\mu\sigma_1}g_{\tau\rho_1}) ] } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{ f^{bs_2t_2}f^{cs_3t_3} G_{\nu\rho_2\sigma_2}(k_2,-k_2-p,p) G_{\lambda\rho_3\sigma_3}(k_3,-k_2-k_3-p,k_2+p) } \nonumber \\ {}~~~~~~~~~~~~~ \times \displaystyle{ g^{\rho_1\sigma_2} g^{\rho_2\sigma_3} g^{\rho_3\sigma_1} \Bigl(\frac{-i}{p^2}\Bigr)^{i+1} \Bigl(\frac{-i}{(p+k_2)^2}\Bigr)^{j+1} \Bigl(\frac{-i}{(p+k_2+k_3)^2}\Bigr)^{l+1} } \nonumber \\ {}~~~~~~~~~~~~~ +\{(\tau,d,k_4) \leftrightarrow (\lambda,c,k_3)\}+\{(\tau,d,k_4) \leftrightarrow (\nu,b,k_2)\} \nonumber \\ {}~~~~~~~~~~~~~ +\{(\mu,a,k_1) \leftrightarrow (\lambda,c,k_3)\}+\{(\mu,a,k_1) \leftrightarrow (\nu,b,k_2)\} \nonumber \\ {}~~~~~~~~~~~~~ +\{(\mu,a,k_1) \leftrightarrow (\nu,b,k_2),~(\tau,d,k_4) \leftrightarrow (\lambda,c,k_3)\}~, \end{array} \label{4g-c} \end{equation} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{equation} \begin{array}{l} \displaystyle{ \Gamma_{(D)\mu\nu\lambda\tau}^{~~~~abcd}(k_1,~k_2,~k_3;q;\mu) ~ = ~ -g_c^{4-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q Tr({\bf F}^a{\bf F}^b{\bf F}^c{\bf F}^d) \frac{1}{N_q}\sum_{i=0}^{q} \sum_{j=0}^{q-i} \sum_{l=0}^{q-i-j} } \nonumber \\ {}~~~~~~~~~~~~ \times \displaystyle{ \int\frac{d^4p}{(2\pi)^4} p_{\mu}(p+k_2)_{\nu}(p+k_2+k_3)_{\lambda}(p-k_1)_{\tau} \Bigl(\frac{i}{p^2}\Bigr)^{i+1} } \nonumber\\ {}~~~~~~~~~~~~ \times \displaystyle{ \Bigl(\frac{i}{(p+k_2)^2}\Bigr)^{j+1} \Bigl(\frac{i}{(p+k_2+k_3)^2}\Bigr)^{l+1} \Bigl(\frac{i}{(p-k_1)^2}\Bigr)^{q-i-j-l+1} } \nonumber \\ {}~~~~~~~~~~~~ +\{(\nu,b,k_2) \rightarrow (\lambda,c,k_3),~(\lambda,c,k_3) \rightarrow (\tau,d,k_4),~(\tau,d,k_4) \rightarrow (\nu,b,k_2)\} \nonumber \\ {}~~~~~~~~~~~~ +\{(\nu,b,k_2) \rightarrow (\tau,d,k_4)),~(\tau,d,k_4) \rightarrow (\lambda,c,k_3),~(\lambda,c,k_3) \rightarrow (\nu,b,k_2)\} \nonumber \\ {}~~~~~~~~~~~~ +\{(\nu,b,k_2) \leftrightarrow (\lambda,c,k_3)\}+\{(\nu,b,k_2) \leftrightarrow (\tau,d,k_4)\}+\{(\lambda,c,k_3) \leftrightarrow (\tau,d,k_4)\}, \end{array} \label{4g-d} \end{equation} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{equation} \begin{array}{l} \displaystyle{ \Gamma_{(E)\mu\nu\lambda\tau}^{~~~~abcd}(k_1,~k_2,~k_3;q;\mu) ~ = ~ -g_c^{4}\mu^{2q} (-i\lambda_q)^{2q} Tr({\bf T}^a{\bf T}^b{\bf T}^c{\bf T}^d) \frac{1}{N_q}\sum_{i=0}^{2q} \sum_{j=0}^{2q-i} \sum_{l=0}^{2q-i-j} } \nonumber\\ {}~~~~~~~~~~~~ \times \displaystyle{ \int\frac{d^4p}{(2\pi)^4} Tr[\gamma_{\mu}\Bigl(\frac{i}{\not{p}-\not{k}_1-m}\Bigr)^{i+1} \gamma_{\tau}\Bigl(\frac{i}{\not{p}+\not{k}_2+\not{k}_3-m}\Bigr)^{j+1} } \nonumber\\ {}~~~~~~~~~~~~ \times \displaystyle{ \gamma_{\lambda}\Bigl(\frac{i}{\not{p}+\not{k}_2-m}\Bigr)^{l+1} \gamma_{\nu}\Bigl(\frac{i}{\not{p}-m}\Bigr)^{2q-i-j-l+1}] } \nonumber \\ {}~~~~~~~~~~~~ +\{(\nu,b,k_2) \rightarrow (\lambda,c,k_3),~(\lambda,c,k_3) \rightarrow (\tau,d,k_4),~(\tau,d,k_4) \rightarrow (\nu,b,k_2)\} \nonumber \\ {}~~~~~~~~~~~~ +\{(\nu,b,k_2) \rightarrow (\tau,d,k_4)),~(\tau,d,k_4) \rightarrow (\lambda,c,k_3),~(\lambda,c,k_3) \rightarrow (\nu,b,k_2)\} \nonumber \\ {}~~~~~~~~~~~~ +\{(\nu,b,k_2) \leftrightarrow (\lambda,c,k_3)\}+\{(\nu,b,k_2) \leftrightarrow (\tau,d,k_4)\}+\{(\lambda,c,k_3) \leftrightarrow (\tau,d,k_4)\}~, \end{array} \label{4g-e} \end{equation} \addtocounter{abc}{1} \renewcommand{\theequation}{\arabic{equation}} \bigskip \noindent 4. \hspace{0.3cm}The integral expressions of the regularized diagram contributing to $\tilde{\Pi}^{ab}(p)$: \begin{eqnarray} \tilde{\Pi}^{ab}(p;q;\mu) & = & -g_c^{2-2q} \mu^{2q}[-ig_c^2C_2({\bf 8})]^q Tr({\bf F}^a{\bf F}^b) \int\frac{d^4k}{(2\pi)^4} p\cdot k \Bigl(\frac{i}{k^2}\Bigr)^{q+1} \Bigl(\frac{-i}{(k-p)^2}\Bigr)~, \label{ghse} \end{eqnarray} \bigskip \noindent 5. \hspace{0.3cm}The integral expressions of the regularized diagrams contributing to $\tilde{\Gamma}^{abc}_{\mu}(p^{\prime},p)$: \renewcommand{\theequation}{\arabic{equation}\alph{abc}} \setcounter{abc}{0} \addtocounter{abc}{1} \begin{eqnarray} \tilde{\Gamma}^{~abc}_{(A)\mu}(p^{\prime},p;q;\mu) & = & ig_c^{3-2q} \mu^{2q}[-ig_c^2C_2({\bf 8})]^q Tr({\bf F}^a{\bf F}^b{\bf F}^c) \int\frac{d^4k}{(2\pi)^4} k^{\sigma}p^{\prime\rho} \nonumber \\ & & \times G_{\mu\rho\sigma}(k_1-k_2,p-k_1,k_2-p) \Bigl(\frac{i}{k^2}\Bigr)^{q+1} \Bigl(\frac{-i}{(k-p^{\prime})^2}\Bigr)\Bigl(\frac{-i}{(k-p)^2}\Bigr)~, \label{glgh-a} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{eqnarray} \tilde{\Gamma}^{~abc}_{(B)\mu}(p^{\prime},p;q;\mu) & = & ig_c^{3-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q Tr({\bf F}^a{\bf F}^b{\bf F}^c) \displaystyle{\int\frac{d^4k}{(2\pi)^4}} (k+p^{\prime})_{\mu}(k+p)\cdot p^{\prime} \nonumber \\ & & \times \displaystyle{\Bigl(\frac{-i}{k^2}\Bigr)^{q+1} \Bigl(\frac{i}{(k+p^{\prime})^2}\Bigr)\Bigl(\frac{i}{(k+p)^2}\Bigr)}~, \label{glgh-b} \end{eqnarray} \addtocounter{abc}{1} \renewcommand{\theequation}{\arabic{equation}} \bigskip \noindent 6. \hspace{0.3cm}The integral expressions of the regularized diagram contributing to $\Sigma(p)$: \begin{eqnarray} \Sigma(p;q;\mu) & = & {\bf T}^a{\bf T}^a \Sigma^{QED}(p;q;\mu) \nonumber \\ & = & -g_c^{2-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q {\bf T}^a{\bf T}^a \int \frac{d^4 k}{(2\pi)^4} (\gamma^{\mu}\frac{i}{ \not{k}-m}\gamma_{\mu}) \Bigl(\frac{-i}{(k-p)^2}\Bigr)^{q+1}, \label{qse} \end{eqnarray} \bigskip \noindent 7. \hspace{0.3cm}The integral expressions of the regularized diagrams contributing to $\Gamma^{a}_{\mu}(p^{\prime},p)$: \renewcommand{\theequation}{\arabic{equation}\alph{abc}} \setcounter{abc}{0} \addtocounter{abc}{1} \begin{eqnarray} \Gamma^{~~~a}_{(A)\mu}(p^{\prime},p;q;\mu) & = & -g_c^{3-2q} \mu^{2q} [-6ig_c^2C_2({\bf 8})]^q {\bf T}^s{\bf T}^t \frac{1}{N_q}\sum_{i=0}^{q} \int\frac{d^4k}{(2\pi)^4} \nonumber \\ & & \times f^{ast} G_{\mu\rho\sigma}(p^{\prime}-p,p-k,k-p^{\prime}) \nonumber \\ & & \times (\gamma^{\rho}\frac{i}{ \not{k}-m}\gamma^{\sigma}) \Bigl(\frac{-i}{(k-p)^2}\Bigr)^{i+1}\Bigl(\frac{-i}{(k-p^{\prime})^2}\Bigr)^{q-i+1}, \label{glq-a} \end{eqnarray} \addtocounter{equation}{-1} \addtocounter{abc}{1} \begin{eqnarray} \Gamma^{~~~a}_{(B)\mu}(p^{\prime},p;q;\mu) & = & {\bf T}^b{\bf T}^a{\bf T}^b \Gamma^{QED}(p^{\prime},p;q;\mu) \nonumber \\ & = & -ig_c^{3-2q} \mu^{2q}[-6ig_c^2C_2({\bf 8})]^q {\bf T}^b{\bf T}^a{\bf T}^b \int \frac{d^4 k}{(2\pi)^4} \nonumber \\ & & \times (\gamma^{\rho}\frac{i}{\not{k}+\not{p^{\prime}}-m} \gamma_{\mu}\frac{1}{\not{k}+\not{p}-m} \gamma_{\rho}) \Bigl(\frac{-i}{k^2}\Bigr)^{q+1}, \label{glq-b} \end{eqnarray} \addtocounter{abc}{1} \renewcommand{\theequation}{\arabic{equation}} \bigskip \bigskip {\raggedright{\Large \bf Acknowledgment}} \vskip 0.5cm \noindent {The work is supported in part by the National Natural Science Foundation of China. One of the author (YC) is also supported in part by Local Natural Science Foundation of Xinjiang.} \newpage \parindent 0 pt
\section{Introduction} The nature of the glass transition is still poorly understood \cite{mode,glass}. Under slow cooling real glasses reach a metastable phase of free energy larger than that of the crystal phase. Glasses show a strong slowing down of the dynamics when the temperature is lowered and the transport coefficients increase by several orders of magnitude in a narrow range of temperatures. It is natural to think that the appareance of high free-energy barriers is the mechanism responsible for the glass transition. But free-energy barriers are composed of energy barriers and entropy barriers. The question about the relevance of both kind of barriers in real glasses is of the outmost importance. Activated jumps of energy-barriers are strongly dependent on temperature. The typical time $\tau$ to overcome an energy barrier $\De E$ is $\tau\sim\exp(\frac{\De E}{T})$ where $T$ is the temperature. This typical time diverges when the temperature $T$ goes to zero. Conversely, relaxation times related to entropy barriers do not depend directly on the temperature. The simplest way to visualize entropy barriers is the following, Consider a dynamics in which at each time step the system can reach a new state with uniform probability; the typical time to decrease the energy of one unity is $\tau\sim \frac{\Omega_i}{\Omega_f}=\exp(\Delta S)$ where $\Delta S$ is the height of the entropy barrier ($\Omega_i$ stands for the initial available volume of phase space and $\Omega_f$ stands for the final volume of phase space with lower energy). If the effect of energy and entropy barriers is combined one expects that entropy barriers should affect the temperature activated relaxation time in its prefactor $\tau\sim \frac{\Omega_i}{\Omega_f}\exp(\frac{\De E}{T})$. According to that, the relaxation time can diverge independently of the temperature if the phase space volume of lower energy configurations in the system shrinks to zero during the dynamical evolution. The idea that an entropy crisis could be relevant to the glassy transition is very old \cite{kauzmann,adam}, and it has had interesting developements in recent times \cite{kirtirwo,par} in the framework of mean-field theory of disordered systems. However in the models studied in \cite{kirtirwo,par} it is very difficult to disentangle entropic effects from energetic ones. To this aim a simple model (the backgammon -BG- model) was recently proposed by one of us \cite{I} (hereafter referred as I), in which energy barriers are completely absent (a diffusive model with entropy barriers has also been considered in \cite{BM}). While the model has no thermodynamic transition, it shows a slow dynamics with strong hysteresis effects and Arrhenius behavior of the relaxation time. The off-equilibrium dynamics of this model was studied subsequently by us \cite{II} (hereafter referred as II) using an adiabatic approximation, obtaining fair good results concerning the relaxation of the energy. The same approximation has been recently rederived, and slightly refined, in a paper by Bouchaud, Godreche and M\'ezard \cite{BMG}. In this paper we derive the exact mean-field equation for the order parameter for the dynamics of the BG model, which turns out to be the energy itself. The techniques we use are similar to these of \cite{BMG}, however, the equations we get were not discussed there. We find that the energy verifies a causal functional equation with memory. This is at variance with the approximate treatments where the evolution is described by a Markovian equation. In its original formulation, the model does not have any energy barriers. However, in real systems energy barriers are present. The BG model is flexible enough to allow for the introduction (and the tuning) of energy barriers. This is done by simple modifications of the Hamiltonian of the system. The formalism developed for the original model applies in these cases. In the second section we define the Hamiltonian of the BG model and the Monte Carlo dynamics we have used to study it. In the third section we present some exact results for the behavior of the one-time quantities (for instance, the energy) and for the two-time quantities like the correlation function. The fourth section is devoted to the study of the effect of energy barriers in the BG model. Finally we present the conclusions and a discussion of the results. \section{The BG model and the dynamics} Let us take $N$ distinguishable particles which can occupy $M$ different states and let us denote by $\rho=\frac{N}{M}$ the density, i.e. the number of particles per state. The BG Hamiltonian is defined by, \be H=-\sum_{r=1}^M\,\de_{n_r,0} \label{eq1} \ee where $n_r$ is the occupation level of the state $r=1,...,M$, i.e. the number of particles which occupy that state. The numbers $n_r$ satisfy the global constraint, \be \sum_{r=1}^M\,n_r= N~~~~. \label{eq2} \ee Eq.(\ref{eq1}) shows that energy is simply given by the number of empty states (with negative sign). We define the occupation probabilities, \be P_k=\frac{1}{M}\,\sum_{r=1}^M\,<\de_{n_r,k}> \label{eq3} \ee which is the probability of finding one state occupied by $k$ particles. The statics of this model in the canonical ensemble can be easily solved (see (I) and (II)). In particular one gets the result, \be P_k=\rho \frac{z^{k-1} \exp(\beta \de_{k,0})}{k!\exp(z)} \label{eq4} \ee where $z$ is the fugacity and $\beta$ is the inverse of the temperature $T$ and they are related by the condition, \be \rho(e^{\beta}-1)=(z-\rho)e^z~~~~. \label{eq5} \ee expressing that the density is fixed to $\rho$. The probabilities $P_k$ satisfy the relation $\sum_{k=0}^{\infty}P_k=1$ and they yield all the static observables, in particular the energy $U=-P_0$. Several dynamical rules, thermalizing to the Boltzmann distribution, can be attached to the model. The simplest choice (I) is the Metropolis single particle dynamics, in which at each sweep a particle is chosen at random and it is proposed a move to a new state. The move is accepted with probability one if the energy does not increase and with probability $\exp(-\beta)$ otherwise. In the mean-field version of the model, the possible arrival states of the particles are chosen at random with uniform probability in all the space. This random motion of the particles allows a complete analytical treatment of the problem \footnote{The interesting case of a sequential dynamics is more complicated.}. Finite dimensional models, where at each sweep the particles are only allowed to move to neighbours on a lattice are currently under study \cite{rito_folla}. The model has no energy barriers. Consequently there is no frustration (in the usual sense) and no metastable states. However it was shown in (I) that the dynamics is highly non trivial and a dramatic slowing down occurs at low temperatures. This can be qualitatively understood as follows. Suppose the system is at zero temperature and the dynamics starts from a random initial configuration of high energy. As the system evolves towards the equilibrium more and more states are progressively emptied and the energy decreases. Because the average number of particles per occupied state increases with time (the total number of particles is conserved) the time needed to empty one more state also increases. The result is that the energy goes extremely slowly to its equilibrium value. The dynamical quantities we are interested in are the time-dependent occupation number probabilities \be P_k(t)=\frac{1}{M}\,\sum_{r=1}^M\,\langle\de_{n_r(t),k}\rangle \label{eq6} \ee ($E(t)=-P_0(t)$) and the two time energy-energy correlation function \cite{I}, \bea C_E(t,s)=& &\frac{\frac{1}{M}\sum_r\delta_{n_r(t),0}\delta_{n_r(s),0} -E(t)E(s)}{-E(s)(1+E(s))} \nonumber\\ \equiv & & {P(n_r(t)=0,n_r(s)=0)-P_0(t)P_0(s)\over P_0(s)[1-P_0(s)]} \,\,\,\,t\ge s \label{eq7} \eea At finite temperature, when $t,s>>t_{eq}\sim \exp(\beta)/\beta^2$ (see reference (II) and also section III.B) this function is time translationally invariant. In the regime in which both times $t,s$ are much less than $t_{eq}$, and at all times at zero temperature, the system is off-equilibrium, time-translation invariance does not hold, and the correlation function displays aging (see \cite{I}). \section{Mean-field equations for the dynamics of the BG model} In this section we derive exact mean-field equations for the Monte Carlo dynamics of the BG model. First we adress the dynamical problem associated to the one-time probability distributions $P_k(t)$. These probabilities generate an infinite hierarchy of Markovian equations which can be closed in terms of the only quantity $P_0(t)$. Then we will study the two-time correlation functions in a similar way. For simplicity, we will restrict all the future computations to the case $\rho=1$ (i.e $M=N$), the generalization to an arbitrary density being very simple. \subsection{Dynamical equations for $P_k(t)$} The purpose of this section is to write the dynamical evolution equations for the probabilities $P_k(t)$ and, in particular, for the internal energy $E(t)=-P_0(t)$. An elementary Monte Carlo move consists in a random selection of one particle (hence, the probability to select a particular departure state $d$ is $n_d/N$ where $n_d$ is the occupation level of that state) and moving it to a randomly selected arrival state $a$ with uniform probability independent of the occupation level $n_a$. One Monte carlo step (our unity of time) consists of $N$ elementary moves. In the elementary move there are several processes which contribute to the variation of $P_k(t)$. In appendix A we write explicitly the balance equations, the result is: \be {{\rm d} P_k(t)\over {\rm d} t}= (k+1)(P_{k+1}-P_{k})+P_{k-1}+ P_0(e^{-\beta}-1)(\de_{k,1} - \de_{k,0} - kP_k + (k+1)P_{k+1}) \label{eq10} \ee where the time index for the probabilities $P_k$ has been omitted. The previous equation holds for $k\ge 0$ with $P_{-1}=0$. In particular for $k=0$ we obtain the equation studied in (II), \be \frac{\partial P_0}{\partial t}=P_1(1-P_0) - e^{-\beta} P_0 (1-P_1) \label{eq11} \ee The hierarchy was closed in (II) assuming fast relaxation on the surfaces of constant energy, and slow variation of the energy itself. In this condition eq.(\ref{eq11}) was solved assuming $P_k(t)={\exp[{\beta(t)\delta_{k,0}-z(t)}]}{z(t)^{k-1}\over k!}$ with $\beta(t)$ and $z(t)$ related at all times by eq.(\ref{eq5}). Here we study the hyerarchy (\ref{eq11}) with the method of the generating function, that we define as \be G(x,t)=\sum_{k=0}^{\infty}\,x^k\,P_k(t) \label{eq12} \ee A similar approach was also used in \cite{BMG} where the adiabatic approximation of (II) was rederived and improved\footnote{The technique of the generating function in the study of the dynamics has also been applied to some mean-field spin glass models \cite{BPPR}.}. {}From the equation (\ref{eq10}) it is easy to check that the $G(x,t)$ satisfies the partial differential equation, \be \frac{\partial G(x,t)}{\partial t}=(x-1)[G(x,t) + \la(t) - (1+\la(t))\frac{\partial G(x,t)}{\partial x}] \label{eq14b} \ee with $\la(t)=P_0(t) (e^{-\beta}-1)$. Eq. (\ref{eq14b}) is a non linear partial differential equation, the nonlinearity is contained in the dependence of $\lambda$ on $P_0(t)=G(0,t)$. The equilibrium solution $G_{eq}(x)$ is easily obtained from equations (\ref{eq4}) and (\ref{eq12}), \be G_{eq}(x)=\frac{e^{\beta}-1+e^{zx}}{z\,e^{z}} \label{eq15b} \ee and one can check that this is consistent with eq.(\ref{eq14b}). The previous partial differential equation can be implicitly solved to get $G(x,t)$ as a functional of $\lambda$. The details are presented in the Appendix B, we give here the result \be G(x,t)={\rm e}^{(x-1)D(t,0)}G_0(1+(x-1)B(t,0))+ (x-1)\int_0^t {\rm d} s \lambda(s) B(t,s) \ {\rm e}^{(x-1)D(t,s)} \label{aa}\ee where we have written \begin{eqnarray} B(t,s)&=&\exp\left({-\int_s^t {\rm d} v \ [1+\lambda(v)]}\right) \nonumber\\ D(t,s)&=&{\int_s^t {\rm d} v \ B(t,v)} \end{eqnarray} and $G_0(x)=G(x,0)$ is the initial condition at time $t=0$. Setting $x=0$ in (\ref{aa}) we get a closed equation for $P_0(t)$ which reads \be P_0(t)= {\rm e}^{-D(t,0)}G_0(1-B(t,0)) +(1-e^{-\beta})\int_0^t {\rm d} s P_0(s) B(t,s) \ {\rm e}^{D(t,s)} \label{bb} \ee The previous equation, although strongly non-markovian is {\it causal}, as the l.h.s. depends on the values of $P_0(s)$ for $s\le t$. It has a unique solution that can be found numerically with good precision, discretizing the time and integrating it step by step. The evaluation of the previous expressions gives the full solution of the BG model as far as the one-time dynamical quantities are concerned. The solution of (\ref{aa}) is explicit at infinite temperature ($\beta=0$). In this case $\la(t)=0$ and the solution of $G(x,t)$ simplifies, \be G(x,t)=e^{(1-e^{-t})(x-1)}\,G_0((x-1)e^{-t}+1) \label{cc} \ee It is not surprising that at infinite temperature the system goes exponentially fast to the equilibrium (with relaxation time equal to 1). At infinite temperature the equilibrium probabilities eq.(\ref{eq4}) are given by $P_k=\frac{1}{k!e}$, the energy being $E=-P_0=-\frac{1}{e}$. If we start from the initial condition in which all particles occupy the same state ($P_0=1,\,P_k=0,k>0$) then we have $G_0(x)=1$. From eq.(\ref{cc}) we obtain the time evolution of the energy, \be E(t)=-G(0,t)=-e^{e^{-t}-1} \label{dd} \ee We studied numerically the solution of (\ref{bb}) at $T=0$. In figure 1 we display the result for the energy, starting from the initial condition $P_k(0)=1/(e k!)$ at time 0 (i.e. $G_0(x)=e^{x-1}$). For comparison we plot the results of the Monte Carlo simulations and of the adiabatic hypothesis of (II) with the same initial condition.\footnote{The adiabatic hypothesis gives better results if the integration is started at latter times.} \subsection{The Correlation Function $C_E(t,s)$} In this section we investigate the behavior of the energy-energy correlation functions (\ref{eq7}). We proceed in a similar way as we have done for the occupation probabilities. We need to study the joint occupation probability in a given site $r$ at two different times $t,s$ ($t>s$), $P(n_r(t)=0,n_r(s)=0)=P(n_r(t)=0|n_r(s)=0)P_0(s)$. The correlation function eq. (\ref{eq7}) can be written as \be C_E(t,s)={P(n_r(t)=0|n_r(s)= 0)-P_0(t)\over 1- P_0(s)}. \ee We now write a set of equations that allow to study $P(n_r(t)=0|n_r(s)= 0)$. Let us define the probabilities \be \nu_k(t,s)=P(n_r(t)=k|n_r(s)=0) \ee i.e. the occupation number probabilities in the set $S_s$ of states which are empty at time $s$. In general, it is possible to restrict the balance equations that led to (\ref{eq10}) to any subset $S$ of the whole space. Irrespectively of $S$ the result is: \begin{eqnarray} \frac{\partial \nu_k}{\partial t}= & &\nu_{k-1}-\nu_{k}+[(k+1)\nu_{k+1}-k\nu_k][1-P_0(1-e^{-\beta})] \nonumber\\ & &-(\de_{k,1}-\delta_{k,0})[\nu_0(1-P_1)+\nu_1 P_0](1-e^{-\beta}). \label{nuk} \end{eqnarray} In particular if the set $S$ is the whole space $\nu_k=P_k$ and we get back to (\ref{eq10}). Of course the initial conditions depend on the set under study. For the set $S_s$ we are interested to, we must choose \be \nu_k(s,s)=\delta_{k,0}. \ee In terms of the generating function \be \G(x,t,s)=\sum_{k=0}^\infty x^k \nu_k(t,s) \ee eq.(\ref{nuk}) reads \be \frac{\partial \G}{ \partial t}= (x-1)[\G-(1-P_0(1-e^{-\beta}))\frac{\partial \G}{ \partial x} -(\nu_0(1-P_1)+\nu_1 P_0)(1-e^{-\beta})] \label{gamma}\ee with condition at time $s$ \be \G_s(x)\equiv \G(x,s,s)=1. \ee Note that if we suppose to know the $P_k(t)$ then the eq.(\ref{nuk},\ref{gamma}) are linear. Obviously, if one considers the set $\overline{S}_s$ complementary to $S_s$, and its respective generating function $\overline{\Gamma}$, it holds the equality: $P_0(s) \G(x,t,s)+(1-P_0(s))\overline{\G}(x,t,s)=G(x,t,s)$. Defining \begin{eqnarray} & &\mu(t,s)=[\nu_0(t,s)(1-P_1(t))+\nu_1(t,s)P_0(t)](1-{\rm e}^{-\beta}) \nonumber\\ & &B(t,s)=\exp\left({-\int_s^t {\rm d} v \ (1-P_0(t)(1-{\rm e}^{-\beta})) }\right) \nonumber\\ & &D(t,s)=\int_s^t {\rm d} v \ B(t,v) \end{eqnarray} we find \be \G(x,t,s)= {\rm e}^{(x-1)D(t,s)}- (x-1)\int_s^t {\rm d} u \mu(u,s) B(t,u) \ {\rm e}^{(x-1)D(t,u)} \label{bbb}\ee which depends implicitly on $\nu_0$ and $\nu_1$. In order to find a closed system we have to consider eq.(\ref{bbb}) and its $x$-derivative in $x=0$ \begin{eqnarray} & &\nu_0(t,s)=1+ \int_s^t {\rm d} u \ \left[ -\nu_0(u,s)[1-(1-P_1(u))(1-{\rm e}^{-\beta})]+\nu_1(u,s)\right] \nonumber\\ & & \nu_1(t,s)= \int_s^t {\rm d} u \ \left[ \mu(u,s) B(t,u){\rm e}^{-D(t,u)}(D(t,u)-1) \right] +{\rm e}^{-D(t,s)}D(t,s) \label{volt} \end{eqnarray} The system (\ref{volt}), if one assumes known the probabilities $P_k(t)$, consists in a vectorial linear Volterra equation of second kind for $\nu_0$ and $\nu_1$ which can in all generality be integrated numerically, and in some particular case also analytically. The simplest case is the equilibrium at finite temperature. In that case, $P_k\equiv P_k^{eq}$ and the various functions appearing in (\ref{volt}) are time traslation invariant. In these conditions eq.(\ref{volt}) can be solved in Laplace transform. Simple algebra, and the formula (see e.g. \cite{tricomi}) \be \int_0^\infty d t \ exp(-a \exp(t)-E)=a^{-E} \ \gamma(p,a) \ee ($\gamma$ is the incomplete gamma function), shows that $\nu_0(E)$, the Laplace transform of $\nu_0(t-s)$, is given by \be \nu_0(E)={A(E)+{z-1\over z}(1-E A(E))\over 1-\left[{(z-1)e^z\over (z-1)e^z +1} +{z-1\over z}E\right](1-E A(E))} \label{lap} \ee where we have expressed all the equilibrium quantities in terms of the fugacity $z$ (see eq. (\ref{eq5})), $E$ is the Laplace variable conjugated to time, and \be A(E)={1\over e^z z^{zE-1}}\int_0^z u^{zE-1}e^u. \ee $\nu_0(E)$, as it should, has a pole in $E=0$ with residue $P_0$ coresponding to $\nu_0(t)\to P_0$ for large time. Poles on the real negative $E$ axis correspond to exponential relaxation modes. The largest relaxation time is given by minus the inverse of the value of $E$ in the pole closest to the origin. This can be obtained explicitely for large $\beta$, where $z\approx \beta-\log(\beta)$ is large, from the asympotic expantion of $E A(E)$ for small $E$ \be E A(E)\approx e^{-z}+E. \ee The result is simply $E_{pole}\approx -e^{-z}$ and correspondingly $\tau_{max}\approx e^z\approx exp(\beta)/ \beta$. In the off-equilibrium regime the integration of (\ref{volt}) can be performed numerically. In fig. 2 we show the result of the integration for $T=0$ for different values of $s$ ( i.e. different waiting times) compared to the Monte Carlo results. Although we did not try very sophisticated algorithms, with standard ones \cite{num_rec}, we were able to reach enough large times to see the scaling behavior $C_E(t,s)=f((t-s)/s)$ observed numerically in I. It would be interesting to see if equation (\ref{bbb}) could be solved with the aid of some simple approximation as it is the case for the energy (II and \cite{BMG}). \section{The effect of energy barriers} The BG model has no energy barriers and hence there is no finite temperature thermodynamic phase transition. In real glasses energy barriers are always present and it can be instructive to understand their effect when combined with entropy barriers. One can easily modify the Hamiltonian (\ref{eq1}) to include energy barriers. In this paper we have focused on two different ways. In the first, we have considered interaction between the different states, introducing an energy gain when groups of states are simultaneously empty. This interaction term is enough to make appear a finite temperature thermodynamic transition, but metastability and frustration are absent and the system monotonically reaches the ground state at zero temperature. In the other, we introduced metastable configurations in the dynamics. In this case the system fails to reach the ground state at zero temperature while thermodinamically there is no finite-temperature phase transition. \subsection{The $p$-states model} The simplest way we can introduce interaction among different states in the model is the following, consider the quantity \be M[\{ n_r\}] ={1\over N}\sum_{r=1}^N (\delta_{n_r,0}-1/e). \ee Any Hamiltonian of the form \be H=N F(M[\{ n_r\}]) \label{hf} \ee with $F$ gentle enough, is a good candidate for a mean-field model. We did not try a systematic study of the form (\ref{hf}) for generic $F$, but we concentrated to the class of monomials, where \be H_p=-\frac{1}{N^{p-1}}\left(\sum_{r=1}^N(\de_{n_r,0}-{1}/{e})\right)^p \label{23} \ee For $p=1$ this model reduces to the BG model. For larger values of $p$ there is interaction between different states. The ground state of this model is the same as the one of the BG model (all particles occupying the same state) and there are no energy barriers at zero temperature. A careful study of the thermodynamics of this model shows that for any $p>1$ there is a first order phase transition from a completely disordered phase with $M=0$ for $T>T_c$ to an `ordered' phase with $M\ne 0$ for $T<T_c$. This leads to the curious situation that the completely disordered state is dynamically stable at all temperatures but at $T=0$. This can be understood by a simple argument. Suppose to start the dynamics in a random initial condition and consider a sweep that lead to the filling of an empty state. The energy variation in this process is \be \delta H=-{1\over N^{p-1}} \left[ \left(\sum_r(\delta_{n_r,0}-{1\over e})-1 \right)^p -\left(\sum_r(\delta_{n_r,0}-{1\over e}) \right)^p \right]\simeq {1\over N^{p-1}} p \left(\sum_r(\delta_{n_r,0}-{1\over e}) \right)^{p-1}. \ee But according to our hypothesis $\sum_r(\delta_{n_r,0}-{1\over e})$ is a random variable of order $\sqrt{N}$, correspondingly $\delta H\sim N^{-(p-1)/2}$ and the acceptance rate \be {\rm e}^{-\beta \delta H}\sim {\rm e}^{-\beta /N^{(p-1)/2}} \ee For finite temperature and large $N$ all the moves are accepted and the energy in average never decreases. In other words the statistic of configurations is not changed by the dynamics. A crossover is found for $\beta\sim N^{(p-1)/2}$ showing that the relevant scale of temperature for the dynamics is different from that of the statics. Right at zero temperature, where only the sign of the energy change and not the magnitude matters, the dynamics of the model coincides for any $p$ with the one of the conventional case $p=1$. \subsection{The effect of metastability} The $p$-states model has no metastability at zero temperature. We want to study here a simple model where metastability is present but without interaction. In the BG model the ground state is reached by emptying progressively more and more states. To empty a given state at a certain time $t$ it is necessary to pass in a configuration where a unique particle occupies that state. We then consider the following model, \be H=\sum_{r=1}^N\,(-\de_{n_r,0}\,+\,g\de_{n_r,1}) \label{30} \ee where $g$ is positive constant and we have the usual constraint eq.(\ref{eq2}). At zero temperature the transition $n_r=2 \ \to \ n_r=1$ is forbidden, hence energy barriers are present in the model. More general models are obtained including in the Hamiltonian all possible terms of the type $\de_{n_r,k}$ \be H=-\sum_{r=1}^N\,\sum_{k=0}^{\infty}g_k\de_{n_r,k}=-\sum_{r=1}^N\,g_{n_r} \label{31} \ee We focus here to the case (\ref{30}). The statics of this model is easily solved. We obtain the free energy, \be \beta f=\log(z)-\log(e^z+e^{\beta}-1+z(e^{-\beta g}-1)) \label{32} \ee and the fugacity is related to the temperature $\frac{1}{\beta}$ by the $g$-independent relation eq.(\ref{eq5}). The equilibrium probabilities $P_k$ (see eq.(\ref{eq3})) are given by, \be P_k=\frac{z^{k-1} \exp(\beta\de_{k,0}-\beta g \de_{k,1})}{k! (e^z+e^{-\beta g}-1)} \label{33} \ee The dynamics of this model is expected to be substantially different to that of the BG model at least at very low temperatures. Concretely, at zero temperature the ground state is the same as for the BG case but there is a large number of metastable states (for instance, half states empty and half of the states with two particles). It is easy to show that for each value of $E$ between $E=-1/2$ and the ground state $E=-1$ there exists a metastable configuration with that energy. Then we expect the value of the energy extrapolated to infinite time to depend strongly on the initial configuration. In order to minimize the energy we have to maximize $P_0$ and minimize $P_1$. While the maximization of $P_0$ is a process where entropy barriers are dominant (this is the reason why the BG model defined as $E=-P_0$ is interesting) this is not the case for minimizing $P_1$ where entropy barriers are absent. Then, independently of the initial configuration we expect that $P_1$ will go to zero in the exponentially fast for large times. In these conditions, we do not expect that the adiabatic solution of (II) can give a good approximation of the dynamics. This approximation was based on the fact that in the BG the surfaces of constant energy are connected, a situation which does not hold here. However the dynamics of this model can be directly solved as in the BG case. Skipping all the details we find for the generating function: \bea \frac{\partial G(x,t)}{\partial t}=(x-1)\Bigl(-(1+\la(t))\frac{\partial G(x,t)}{\partial x} + (1+2 P_2(e^{-\beta g}-1))G(x,t)+\\ \la(t)-2 P_2 (e^{-\beta g}-1)(e^x-P_0 (x-1)(1-e^{-\beta (1+g)})\Bigr ) \label{36} \eea where $\la(t)=P_0(t) (e^{-\beta(1+g)}-1)$. Observe that eq.(\ref{36}) depends only on the probabilities $P_0(t)$ and $P_2(t)$. The solution is more complicated to that of eq.(\ref{aa}), however it can be found. In figure 3 we show the result of the numerical integration of (\ref{36}) for the energy at $T=0$ compared with the Monte Carlo simulations, starting from a completely random initial condition. The energy seems to converge to a value $\lim_{t\to\infty}E(t)= -.564$, a result that it would be interesting to derive analytically. It is under current study the finite temperature dynamics, where we expect the effect of the energy and entropy barriers to combine to give rise to a dynamics slower than that of the BG model. \section{Conclusions} In this paper we have derived the exact mean-field equations of the dynamics of the Backgammon model. This has been achieved through the study of the single site occupation number probability for which a hierarchical set of equations hold. With the method of the characteristic function, we have derived a closed functional equation for the energy. This, although non-Markovian, has a causal character and can be integrated step by step discretizing the time. The non-Markovian character of the evolution equation, suggests that history dependent effects should be observable in the system. However the analysis of II, where the evolution of the energy was described by an approximate equation, shows that even in subtle phenomena as hysteresis cycles in cooling-heating processes, history dependent effects are very small. This should reflect in the fact that the memory kernels that appear in the equation for the energy are short range in time. The method of the generating function also allowed us to derive also a system of linear Volterra equations describing the evolution of the energy autocorrelation function. The numerical solution of these equation confirmed the aging behaviour found in I. It would be interesting to derive analytically the scaling $C_E(t,t')=f(t'/t)$. In the last section we have derived the mean-field theory for a model where entropic and energetic barriers are combined. We have seen that at temperature zero, starting from a random configuration the system fails in finding the ground state. For future work it is left the study of this model for finite temperature. Non linear equations with memory appear in phenomenological glass theory under the name of Mode Coupling Theory \cite{mode}. Mode coupling equations appear naturally in the mean-field treatment of the dynamics of disordered \cite{cuku,frame} or quasi-disorderd systems \cite{FH}, in off-equilibrium situation they involve a set of coupled integral equation for the two time auto-correlation function and its conjugated response function. The most striking manifestation of the importance of memory effects in off-equilibrium mode coupling theory is in the aging behaviour of the response function \cite{cuku,frame}. Structural glasses are generally classified as strong glasses (Arrhenius behavior of the relaxation time) or fragile glasses (Vogel-Tamman-Fulcher behavior of the relaxation time). In this classification the BG model is a strong glass. Polymer glasses are fragile glasses which show strong aging effects in its physical properties \cite{glass}. It would be desiderable to know from experiments if there is a correlation between the fragility of glasses and its aging properties. This could shed light on the role of energy barriers in the mechanism of the glass transition. We believe that only entropy barriers cannot yield aging effects in the response function. In this framework a more detailed study of the BG model with metastability (as presented in the last section) at finite temperature could be instructive. In particular, the study of the relaxation time as a function of the temperature and the existence of aging due to the presence of energy barriers. \begin{center} {\bf Acknowledgements} \end{center} We would like to thank J.P. Bouchaud, E.Follana and M. M\'ezard for interesting discussions. (F.R.) acknowledges Ministerio de Educacion y Ciencia of Spain for financial support. \section{Appendix A} In this appendix we derive the evolution equation for the probability $P_k(t)$. Define as $N_k(t)$ the number of states occupied by $k$ particles ($P_k(t)=N_k(t)/N$). The processes leading to a variation of $N_k$ can be classified in this way: \begin{itemize} \item{{\it Process A$+$}: arrival of a particle in a state with $k-1$ particles.} \item{{\it Process A$-$}: departure of a particle from a state with $k$ particles.} \item{{\it Process B$+$}: departure of a particle from a state with $k+1$ particles.} \item{{\it Process B$-$}: arrival of a particle in a state with $k$ particles.} \end{itemize} Note that in a sweep the processes above are not mutually exclusive, so, for example the contemporary occurency of $A+$ and $B-$ lead to no variation in $N_k$. At each Monte Carlo sweep three independent random variables are extracted: a departure state $d$ with probability $n_d/N$, an arrival state $a$ with probability $1/N$ and an acceptance variable \be x= \left\{ \begin{array}{lll} 1 & {\rm with \;\;\; prob.} & {\rm e}^{-\beta} \\ 0 &{\rm with \;\;\; prob.}&1- {\rm e}^{-\beta} \end{array} \right. \ee In terms of these variables the variation in $N_k$ in each process is given by: \begin{itemize} \item{{\it Process A$+$}: $\ \ \de_{n_a,k-1} [1-\de_{k,1}(1-\de_{n_d,1})\delta_{x,0}]$} \item{{\it Process A$-$}: $\ \ -\de_{n_d,k}[1-\de_{n_a,0}\delta_{x,0} + \de_{k,1}+\de_{k,1}\de_{n_a,0}\delta_{x,0}]$} \item{{\it Process B$+$}: $\ \ \de_{n_d,k+1}[1-\de_{n_a,0}\delta_{x,0} +\de_{k,0}\de_{n_a,0}\delta_{x,0}]$} \item{{\it Process B$-$}: $\ \ -\de_{n_a,k} [1-\de_{k,0}(1-\delta_{n_d,1})\delta_{x,0}] $} \end{itemize} The contribution in the different processes can be easily understood, for example in process A$+$ we must have $k-1$ particles in the arrival state. If $k=1$ and $n_d>1$ the move imply an energy cost, and is accepted only if $x=1$. Summing all the contribution and averaging over $p$, $a$ and $x$ we find: \bea & & \lan N_k(t+\delta t)-N_k(t)\ran = N[P_k(t+\delta t)-P_k(t)]\equiv {{\rm d} P_k(t)\over {\rm d} t}= \nonumber\\ & &(k+1)(P_{k+1}-P_{k})+P_{k-1}+ P_0(e^{-\beta}-1)(\de_{k,1} - \de_{k,0} - kP_k + (k+1)P_{k+1}) \label{A10} \eea Very similar considerations lead to (\ref{nuk}) if one restricts the balance equation to a subset of the whole space. \section{Appendix B} In this Appendix we obtain the solution of eq.(\ref{eq14b}). We perform the change of variables $(x,t)\to(u,t)$ where $x-1=e^{u+\int_0^t{\rm d} s (1+\lambda(s))} =e^u B(t,0)$. In terms of the new variables eq.(\ref{eq14b}) reads, \be \frac{\partial \hat{G}}{\partial t}= e^u B(t,0)(\hat{G} +\la) \label{ap2} \ee where $\hat{G}(u,t)=G(x(u,t),t)$ This is a linear differential equation which can be readily solved \be \hat{G}(u,t)=e^{e^u\int_0^t\, {\rm d} s B(s,0)} F(u) + e^u \int_0^t {\rm d} s \ \lambda(s)B(s,0)e^{e^u\int_s^t\, {\rm d} v B(v,0)} \label{ap3}\ee where $F$ is an arbitrary function. Going back to $(x,t)$ and imposing the initial condition we get eq.(\ref{aa}).
\section{Introduction} Reduced-dimensionality structures are currently attracting much attention.\cite{P5,M1,K2,P6} Modern technology made it possible to create nanoscale 0D structures where motion in all directions is quantized (quantum dots). Often a system represents a large array of independent quantum dots. Because of extremely small dimensions, the fluctuations of parameters of individual quantum dots become an important factor since they determine the properties of the whole array. This paper was initiated by the experimental work Ref.\ \onlinecite{P5}, where nice photoluminescence (PL) and photoluminescence excitation (PLE) experiments were performed with an array of self-assembling quantum dots. The conventional nonselective (with above-barrier excitation) PL reveals a broad peak of about $\sim$ 50 meV in width. This width originates from a wide spread of energies of different quantum dots in the array. Both PLE and selectively-excited PL spectra (when only the quantum dots that are in resonance with incident light are excited) show sets of broadened peaks with 2-3 times smaller widths. These peaks correspond to the distribution of energy levels in the subset of quantum dots that are resonantly excited. It was observed that the first PLE peak is strongly shifted from the detection energy. The origin of this shift remained unclear. Moreover, no measurable Stokes shift was observed in a recent paper Ref. \onlinecite{K2}, where the PL has been measured together with the direct absorption by the layer of quantum dots. In this paper a simple theory of the PL from an array of quantum dots is developed. We suggest that the process of photoexcitation of a dot into its lowest optically-excited state does not contribute to the PL signal. The proposed interpretation of the experimental data explains large Stokes-like shift between the PL and PLE peaks as the average distance between the two lowest optically-excited states. In what follows we use the term ``ground state'' to denote the lowest optically-excited state because it corresponds to the ground state of the photoexcited carriers in the dot. It is suggested that the PL and PLE lineshapes are completely determined by the statistical distribution of the energy levels of different quantum dots in the array. More precisely, it is determined by the distribution of {\em pairs} of energy levels. It is shown that such distribution is essentially two-variable. That is, there exists a correlation in the positions of different energy levels in each quantum dot throughout the array, however, such correlation is not 100\%. We show that this feature causes the difference in the positions of the maxima of the PL and PLE spectra. The fluctuations of energy levels due to the random potential caused by the alloy composition fluctuations are studied in Section IV. We show that the major part of the observed linewidth can be accounted for by this mechanism. We also suggest that the first two excited states observed in Ref.\ \onlinecite{P5} originate from the twofold degenerate first excited level, when the degeneracy is lifted by the random potential. The density of states and the two-level distribution function, which accounts for the correlation in energy-level positions, are calculated. The spectra determined by these functions describe most of the features of the experimental spectra. Finally, the effects caused by fluctuations of the {\em shape} of the quantum dots are discussed. It is suggested that the fluctuations of the shape of the dots should increase the energy-level correlation. \section{Origin of the Stokes-like Shift} In the present paper we consider the case when both the electron and the hole are confined in the quantum dot. Energy of the quantum dot is measured from the unexcited state of the dot with no electrons and holes. The term ``ground state'' refers to the ground state of the quantum dot with an excited electron-hole pair, i.e.\ to the lowest optically-excited state. The key feature of an array of quantum dots that differs it from a quantum well or bulk material is that there is no charge transfer between the dots, or at least such transfer is strongly suppressed. Different quantum dots therefore give independent additive contributions in any optical experiment.\cite{C1} The density of states for each quantum dot represents a set of delta-function-like peaks, while the total density of states of the whole system can be spread in a wide energy range due to the inhomogeneous broadening. Several experiments have been reported recently where the contributions to the luminescence from {\em single} quantum dots were found.\cite{M1,K2,P6,A1} Single quantum dots give extremely narrow sub-meV spectral lines. When, however, the number of the excited quantum dots is large, a broad luminescence peak is observed.\cite{P5,K2,P6} We suggest that such a feature makes it difficult to probe optically the ground states of the dots and thus causes an apparent Stokes-like shift observed in Ref.\ \onlinecite{P5}. Probing of the ground states of the dots may require special technique like high resolution or time-resolved PL. Indeed, let's consider the contribution to the PL from the process when the photon is absorbed into the ground state of a dot and then re-emitted. If the optical process is not phonon-assisted, the energies of the incoming and outgoing photons are exactly equal. The PL response can hardly be observed since it is hidden by the incident beam scattered by other elements of the experimental environment. This situation is quite different from that in quantum wells or bulk semiconductors, where an electron excited to the conduction band has always quantum states with smaller energies. It can loose some energy before it recombines in an {\em independent} optical process. In a quantum dot, the emitted photon may have its energy below (or above) that of the exciting light if the PL is phonon-assisted. However, the phonon must be emitted (absorbed) together with photon {\em in the same quantum process}. The probability of such process is determined in higher-order perturbation theory in the electron-phonon coupling constant and is therefore much smaller than the probability of the direct transition. In order to give a substantial contribution to the PL signal, the dot must be pumped into one of its {\em excited} states. Therefore, the minimum distance between the excitation and detection energies seen in the spectra is equal to the distance between the ground and first excited states of the dot. In fact, no Stokes shift was found between the PL and the direct absorption by a layer of quantum dots measured in Ref.\ \onlinecite{K2}. \section{Positions of peaks in the PL and PLE spectra} In this section the shape of the PL and PLE spectra is described qualitatively. Different peaks observed in the spectra are assigned. It is shown that existence of a random spread of interlevel distances in quantum dots causes substantial deviations of the positions of peaks seen in PLE and selectively-excited PL spectra. We shall assume for simplicity that the energy relaxation in a quantum dot occurs faster than recombination, so that the light is always emitted from the ground state of the dot. Let us first consider a simple model when the distances between different energy levels are the same for all quantum dots, however, there is a wide distribution of energy levels in the array. In other words, let the picture of energy levels be the same in each quantum dot but shifted randomly as a whole. This implies a 100\% correlation in the positions of different energy levels in each dot. If the conventional PL technique is used so that the pumping is performed with energies well above the barriers between the dots, all the dots are excited and emit light at their ground-state energies. The emission then represents a broad peak due to the wide distribution of the ground-state energies. Let us now consider PLE and selectively-excited PL measurements. Here only those dots are excited, which have one of their energy levels in resonance with the exciting light. As suggested above, the dots pumped into their ground states do not contribute to the spectra. When the interlevel distances are the same in all dots, both PLE and selectively-excited PL spectra would show a set of delta-function-like peaks as sketched in the Fig.\ 1. It is more convenient to start with the PLE (left). By fixing the detection energy one selects the subset of all quantum dots in the array with the ground-state energy $E_0=E_{detector}$ (we have assumed that the light is always emitted from the ground state of the dots). The PLE signal appears when the excitation energy matches the energy of an excited state ($E_{laser}=E_1$, $E_2$, etc.) in the selected subset. Thus the first observed (lowest in energy) PLE peak corresponds to the {\em first excited energy levels of such dots} that have their ground-state level at the detection energy. The analysis of the selectively-excited PL (right) is somewhat more complicated but similar. The fixed energy of excitation selects {\em several} subsets of all quantum dots such that $E_1=E_{laser}$, $E_2=E_{laser}$, etc. A nonzero PL signal appears when the detection energy matches the ground-state energy of one of the subsets. The position of the first observed PL peak (highest in energy) thus corresponds to the {\em ground state energy of such dots} that have their {\em first excited} energy level at the excitation energy. The distance between the first two PL peaks is equal to $E_{21}$, the distance between {\em the first and the second} excited states of the dots. When interlevel distances are the same in all dots, all distances between different peaks in the PL (PLE) series also remain the same, while the whole picture shifts with the shift of the laser (detector) energies. It means that positions of these peaks being plotted against the laser (detector) energies must form a set of straight lines with the unit slope. The slopes of such lines obtained from the PLE and selectively-excited PL spectra obtained in Ref.\ \onlinecite{P5} are all less then 1. For the first PL and PLE peaks, e.g., $dE_{max,1}^{PL}/dE_{laser}=0.91$, $dE_{max,1}^{PLE}/ dE_{detector}=0.77$.\cite{P5} It is easy to see that such deviation cannot be attributed to the dependence of the interlevel distances on the ground-state energy. Indeed, in this case the slopes obtained from the PL and PLE spectra should be inverse of each other. We see, however, that \begin{equation} \frac{dE_{max,m}^{PL}}{dE_{laser}}\neq \left(\frac{dE_{max,m}^{PLE}}{dE_{detector}}\right)^{-1}. \end{equation} We show below that existence of a spread of interlevel distances in the array causes deviation of the positions of peaks seen in the PL and PLE spectra in such a way that both derivatives in Eq.\ (1) become less then 1. To demonstrate this, let us assume that the interlevel distances in the dots fluctuate randomly in some (narrow) energy interval. In this case selection of a subset of quantum dots by fixing one of the energy levels does not yet determine the positions of other energy levels. Then one should observe a sequence of broadened peaks in both PL and PLE. The scale of broadening is determined by the distribution of interlevel distances only, therefore it can be narrower than the distribution of the absolute energies. As shown below such behavior is natural if the spread is caused by a random potential. To understand how the {\em positions of maxima of the broadened peaks} are shifted, it is convenient to draw a three-dimensional picture shown in Fig.\ 2. Here the intensity measured by the detector is plotted as a function of two variables --- the excitation and the emission energies. Clearly, such a plot contains all information, which both PL and PLE can provide. To get the shapes of the PL or PLE spectra one simply has to slice the three-dimensional plot along the $E_{laser}$ or $E_{detector}$ axis. First, the intensity is zero in the half-plane $E_{laser}<E_{detector}$. The intensity is nonzero only when $E_{laser}$ and $E_{detector}$ are in resonance with the excited and ground states of some dot {\em simultaneously}. If the spread of interlevel distances is small, the three-dimensional plot has a shape of a set of narrow {\em ridges}, elongated in the direction parallel to the line $E_{laser}=E_{detector}$. Each ridge corresponds to the optical process for which $E_{detector}=E_0$ and $E_{laser}=E_m$, $m$=1, 2, etc. The width of each ridge is determined by the spread of the corresponding interlevel distances, while the length is larger and represents the large spread of the ground-state energies. The distance between the ridge and the line $E_{laser}=E_{detector}$ is determined by the average distance between the ground and the corresponding excited states of the dots. The inset in Fig.\ 2 shows a fragment of the same plot in the isoline projection (dashed lines). Crosses and pluses show the positions of the maxima in the PL and PLE crossections respectively. The ``PL and PLE lines'' are the loci of the maxima observed in the PL and PLE spectra. As one can see, the positions of the maxima deviate from the major axis of the ridge and from each other. Indeed, these positions lie in such points where the crossection line (parallel to one of the axes) is tangential to the lines of equal intensity. Note that the deviation of the positions of the maxima is such that both derivatives in the Eq.\ (1) are less than 1. Both lines intersect exactly at the maximum of the ridge, which coordinates give the average positions of the ground and the excited states of all the dots in the array. In general, if the oscillator strengths are the same for all allowed transitions, the intensity as a function of the excitation and detection energies is proportional to the mutual level-level distribution function: $I(E_{detector},E_{laser})\propto\sum_m P(E_{detector}= E_0,E_{laser}=E_m)$, where $E_m$ are the energy-level positions, $E_0$ being the position of the ground state. In the following section we derive the shape of the distribution function $P(E_0,E_m)$ when the spread of the energy-level positions is caused by the composition fluctuations. We show that a quite complicated PL and PLE lineshape observed in Ref.\ \onlinecite{P5} can be sufficiently well described in this way. \section{fluctuations of energy levels in quantum dots} In this section we study the properties of the statistical distribution of the energy levels in quantum dots caused by a white-noise random potential. We present a strong evidence that the major part of the observed spread of the PL and PLE peaks is caused by universal composition fluctuations in the dots. These fluctuations are a generic property of semiconductor alloys and produce a theoretical limit for unification of quantum dots in an array. We also suggest that the first two excited levels that reveal themselves in the PLE and selectively-excited PL experiments originate {\em from a single doubly degenerate level} when the degeneracy is lifted by the random potential. In semiconductor alloys the lattice sites are occupied randomly with two types of substitutional atoms. We ignore here the correlation between occupation of different sites. The composition $x$ averaged over a small volume always fluctuates. The order of magnitude of the fluctuations is inversely proportional to the square root of the volume over which the averaging is performed. Though usually small, this effect can be important in extremely small structures. In small quantum dots such composition fluctuations cause shifts of energy levels from the positions determined by the average composition $x_0$. The ``local'' composition $x$ varies from dot to dot, and also {\em inside} the dot. According to that, the energy levels fluctuate from dot to dot. Shifts of the energy levels in different quantum dots are independent from each other. Inside a single quantum dot the shifts of different energy levels are correlated. Note, however, that such correlation is not 100\% (as it would be if there was only one fluctuating parameter like the diameter of a dot). In order to calculate the statistical distribution of energy levels caused by composition fluctuations we use the method developed by Efros and Raikh (for a review see Ref.\ \onlinecite{RE}). This method is applicable if the size of the wave function is much larger than the lattice constant. In this case one needs to know only the shape of the wave function in a quantum dot and the slope of the dependence of the gap on composition, $dE_g/dx$, at average composition $x_0$. The result depends on whether the unperturbed energy levels are degenerate or not. Without degeneracy the distribution of energy levels $E_m$ is Gaussian with the standard deviation given by \begin{equation} \sigma_m^2=\overline{\epsilon_m^2}= \gamma\int d^3r\ \psi_m^4({\bf r}), \end{equation} where $\epsilon_m=E_m-\overline E_m$, $\psi_m$ is the wave function (real) corresponding to the energy level $E_m$, and $\gamma$ is given by \begin{equation} \gamma=\left(\frac{dE_g}{dx}\right)^2\frac{x_0(1-x_0)}{N}, \end{equation} where $N$ is the number of lattice sites per unit volume. The {\em covariance} between $\epsilon_m$ and $\epsilon_n$ can be obtained in a similar way: \begin{equation} \overline{\epsilon_m\epsilon_n}=\rho\sigma_m\sigma_n =\gamma\int d^3r\ \psi_m^2({\bf r})\psi_n^2({\bf r}). \end{equation} The coefficient $\rho$, $\rho\leq1$, is called the coefficient of correlation between $E_m$ and $E_n$. To find the shape of the PLE and selectively-excited PL spectra we need the mutual distribution function for $E_0$ and $E_m$. The most general form of the two-variable Gaussian distribution is given by \begin{eqnarray} &G&_2(\epsilon_0,\epsilon_m;\sigma_0,\sigma_m,\rho)= \frac{1}{2\pi\sigma_0\sigma_m\sqrt{1-\rho^2}}\nonumber\\ &&\times\exp\left\{-\frac{1}{2(1-\rho^2)} \left[\frac{\epsilon_0^2}{\sigma_0^2} -2\rho\frac{\epsilon_0\epsilon_m}{\sigma_0\sigma_m} +\frac{\epsilon_m^2}{\sigma_m^2}\right]\right\}, \end{eqnarray} This is just an analytical expression for the shape of the ridge, discussed in the previous section. The ridge is strongly elongated when the parameter $\rho$ is close to 1. It is equal to 1 in the limiting case when an exact relation between the energy-level positions exists, so that the two-variable statistical distribution becomes effectively one-variable. The ratio $\sigma_m/\sigma_0$ determines the orientation of the ridge. The ridge is parallel to the line $E_{laser}=E_{detector}$ when $\sigma_m=\sigma_0$. The situation is, however, more complicated if a degeneracy exists. The random potential shifts the degenerate energy level and lifts the degeneracy. The distribution of energies in the vicinity of an unperturbed degenerate level appears to be not Gaussian. The mutual two-variable distribution function for a transition between $E_0$ and $E_m$ is not Gaussian either. Instead, it has a shape of two close parallel ridges corresponding to each of the split-off energy levels. For simplicity we restrict ourselves to the axially symmetric quantum dots, where each energy level except the ground state is doubly degenerate. In this case it appears to be possible to derive the general form of the distribution function without knowledge of the shape of the wave functions in the quantum dot. For an axially symmetric system two degenerate wave functions with the angular momentum $|m|$ have the form $\psi_{m\pm}({\bf r})=\psi_m(r) e^{\pm im\phi}$, where $\psi_m$ can be made real. The positions of energy levels $E_{m\pm}$, split and shifted by random potential of the particular configuration, can be obtained as the eigenvalues of the secular matrix \begin{equation} \delta H=\left[ \begin{tabular}{cc} $u$ & $x+iy$\\ $x-iy$ & $u$ \end{tabular} \right], \end{equation} where the matrix elements $u$ and $x+iy$ take random values in each quantum dot and are given by: \begin{mathletters} \begin{eqnarray} u&=&\sqrt{\gamma}\int d^3r\ V({\bf r})\psi_m^2({\bf r}),\\ x+iy&=&\sqrt{\gamma}\int d^3r\ V({\bf r})\psi_m^2({\bf r}) e^{2im\phi}. \end{eqnarray} \end{mathletters} Here $V({\bf r})$ is the white-noise random potential with correlator $<V({\bf r})V({\bf r}')>=\gamma\delta({\bf r}-{\bf r}')$. Eigenvalues of $\delta H$ are $\epsilon_\pm=u\pm\sqrt{x^2+y^2}$ (the energy is measured from the unperturbed energy level). It is easy to see that $u$, $x$, and $y$ are {\em independent} Gaussian random variables with $\overline{u^2}=\sigma_m^2$, $\overline{x^2}=\overline{y^2}=\sigma_m^2/2$, $\sigma_m^2$ being given by Eq.\ (2). The density of states in the vicinity of $E_m$ (the distribution function for $\epsilon_m=E_m-\overline E_m$) is given by \begin{eqnarray} P(\epsilon_m)&=&\sum_\pm\int\!\int\!\int dudxdy\ \delta(\epsilon_m-u\mp\sqrt{x^2+y^2})\nonumber\\ &&\times G_1(u;\sigma_m) G_1(x;\frac{\sigma_m}{\sqrt{2}})G_1(y;\frac{\sigma_m}{\sqrt{2}}) \nonumber\\ &=&D_0(\epsilon_m;\sigma_m), \end{eqnarray} where $G_1$ is the standard one-variable Gaussian distribution, $G_1(\epsilon;\sigma)\equiv\sigma^{-1}G_1(\epsilon/\sigma;1)$, $G_1(z;1)=(2\pi)^{-1/2}\times$ $\exp(-z^2/2)$; and we have introduced the notation for the distribution function $D_0(\epsilon;\sigma)\equiv\sigma^{-1}D_0(\epsilon/\sigma;1)$, \begin{equation} D_0(z;1)=\frac{2}{3}\sqrt{\frac{2}{\pi}}e^{-z^2/2}+ \frac{2z}{3\sqrt{3}}e^{-z^2/3}\text{erf}(\frac{z}{\sqrt{6}}). \end{equation} The meaning of the index 0 in $D_0(\epsilon;\sigma)$ will soon become clear. The Eq.\ (9) describes a bell-shaped curve, which determines the density of states and, hence, the {\em absorption spectrum} in the vicinity of the excited level $E_m$. It's shape can be fairly well approximated by a Gaussian with the effective dispersion $\sigma_{eff}=\sigma_m\sqrt{2}$, as shown in Fig.\ 3. In order to find the mutual two-level distribution function that determines the shape of the PL and PLE spectra it is important to account for the correlation between matrix elements of $\delta H$ and the shift of the ground-state $\epsilon_0$. It is useful to note that the expression for the matrix element $u$ is the same as the expression for the shift of a non-degenerate level with the wave function $\psi_m(r)$. The value of $u$ is, therefore, {\em correlated} with $\epsilon_0$ via the same two-variable Gaussian distribution as in Eq.\ (5): $P(\epsilon_0,u)=G_2(\epsilon_0,u;\sigma_0,\sigma_m,\rho)$, with the parameters $\sigma_0$, $\sigma_m$, and $\rho$ given by Eqs.\ (2), (4). The parameters $x$ and $y$ are statistically independent of $\epsilon_0$ and $u$, because the corresponding covariances become zero when the integration over the azimuthal angle $\phi$ is performed. Thus, we obtain: \begin{eqnarray} P(\epsilon_0,\epsilon_m)&=&\sum_\pm\int\!\int\!\int dudxdy\ \delta(\epsilon_m-u\mp\sqrt{x^2+y^2})\nonumber\\ &&\times G_2(\epsilon_0,u;\sigma_0,\sigma_m,\rho) G_1(x;\frac{\sigma_m}{\sqrt{2}})G_1(y;\frac{\sigma_m}{\sqrt{2}}) \nonumber\\ &=&G_1(\epsilon_0;\sigma_0) D_\rho(\epsilon_m-\epsilon_0\rho\frac{\sigma_m}{\sigma_0};\sigma_m), \end{eqnarray} where the function $D_\rho(\epsilon;\sigma)\equiv\sigma^{-1}D_\rho(\epsilon/\sigma;1)$, and \begin{eqnarray} D_\rho(z;1)&=&\frac{2\sqrt{\mu-1}}{\sqrt{\pi}\mu} \exp\left(-\frac{z^2}{\mu-1}\right)\nonumber\\ &+&\frac{2z}{\mu^{3/2}}\exp\left(-\frac{z^2}{\mu}\right) \text{erf}\left[\frac{z}{\sqrt{\mu(\mu-1)}}\right]. \end{eqnarray} Here $\mu=3-2\rho^2$. The function $D_0(\epsilon;\sigma)$ defined by Eq.\ (9) is a particular case of (11) with $\rho=0$. The function $D_\rho(\epsilon_m;\sigma_m)$ gives the distribution of the energy sublevels in the vicinity of $\overline E_m$, when {\em the position of the ground state is fixed}. It is normalized in such a way that $\int d\epsilon\ D_\rho(\epsilon;\sigma)=2$ (according to the twofold degeneracy), and $\int d\epsilon\ \epsilon^2 D_\rho(\epsilon;\sigma)=2\sigma^2(2-\rho^2)$. The function $D_\rho(\epsilon;\sigma)$ is symmetric in $\epsilon$. It has one maximum when $\rho\leq1/\sqrt{2}$ and two maxima, when $\rho>1/\sqrt{2}$. The maxima become more pronounced when $\rho$ tends to 1. The function $P(\epsilon_0,\epsilon_m)$ determined by Eq.\ (10) gives the probability density for a quantum dot to have the ground state at the energy $\overline E_0+\epsilon_0$ and an excited state at the energy $\overline E_m+\epsilon_m$. It is proportional to the intensity measured by the detector at fixed excitation and detection energies. Hence, it determines the shape of both PL and PLE spectra when the proper argument is fixed. The function $P(\epsilon_0,\epsilon_m)$ appears to be not very sensitive to the ratio $\sigma_m/\sigma_0$. It is, however, quite sensitive to the value of the correlation coefficient $\rho$. This function is plotted in Fig.\ 4 for $\sigma_m=\sigma_0$ and two different values of $\rho$. As shown in the next section, it is natural for the coefficient $\rho$ to be close to 1. The value $\rho=0.94$ [Fig.\ 4(a)] gives the best fit to the experimental data. When $\rho$ tends to 1 [Fig.\ 4(b)] the intensity in the dip between two maxima approaches zero. The lineshapes of the spectra described by Eq.\ (10) are discussed in detail in the next section. \section{discussion} We replot the positions of the PL and PLE maxima observed in Ref.\ \onlinecite{P5} on the combined plot in Fig.\ 5 to illustrate that the experimental behavior is in agreement with our consideration (compare with the inset in Fig.\ 2). The dotted line shows unit slope, $E_{laser}=E_{detector}$. Three lines correspond to the positions of two maxima observed in each PL curve (crosses) and one --- in each PLE curve (pluses). Positions of only one PLE maximum for each $E_{detector}$ are shown because the second PLE maximum is seen not clear enough. The constant value of 1290 meV has been subtracted from all energies. Three lines shown in Fig.\ 5 are consistent with the qualitative picture of two close parallel ridges. As suggested above, these ridges correspond to the two optical processes where $E_{detector}=E_0$ and $E_{laser}=E_{1\pm}$; $E_0$ being the ground-state energy and $E_{1\pm}$ --- the energies of the first and second excited states. We use the notation $E_{1\pm}$ instead of $E_{1,2}$ for the two excited states according to the idea that they originate from the twofold-degenerate first excited energy level $E_1$ when the degeneracy is lifted by random potential. This idea is supported by the fact that the ratio of interlevel distances appears to be $(\overline E_{1+}-\overline E_{1-})/(\overline E_{1-}-\overline E_0)\approx 0.6$. For a cylindrical quantum well with infinite potential walls the corresponding ratio is about $(\overline E_{21}/\overline E_{10})=1.3$. The point of intersection of two lines gives the position of the maximum of the first ridge. The difference of 55 meV in its coordinates is nothing but the observed Stokes-like shift between emission and absorption energies. The distance between two ridges measured along the $E_{laser}$ axis gives $\overline E_{1+}-\overline E_{1-}$, the mean splitting of the excited state $E_1$. Using Fig.\ 5 we may conclude that the average distance between the ground and (split) first excited state is about 75 meV, while the average splitting of the excited state is approximately 40 meV. Fig.\ 6 shows the comparison between the presented theory and the experiment. The experimental spectra from Ref.\ \onlinecite{P5} are replotted in Fig.\ 6(a). The theoretical PLE and selectively-excited PL spectra [Fig.\ 6(b)] are obtained with the use of Eq.\ (10) by fixing $\epsilon_0$ or $\epsilon_m$ respectively. Experimental values for the average positions of the energy levels $\overline E_0=1276$ meV and $\overline E_1=1351$ meV are used. The Fig.\ 6(b) is essentially the Fig.\ 4(a), replotted in the wavelength scale for better comparison. The value $\rho=0.94$ gives the best fit for the data from Ref.\ \onlinecite{P5}. It is seen that the curves obtained reproduce all the features of both PL and PLE spectra. The structure of the expression (10) is such that if $\epsilon_0$ is fixed, it describes a curve symmetric around the point $\epsilon_m=\epsilon_0\rho\sigma_m/\sigma_0$. Thus (if $\rho>1/\sqrt{2}$) the PLE line should reveal two {\em symmetric} peaks at any detection energy. This is exactly the behavior seen in the experiment. When the detection energy is changed, the whole spectrum must shift linearly with it. Indeed, Fig.\ 5 shows that the positions of PLE maxima depend linearly on the detection energy. The slope of the PLE curve in Fig.\ 5 gives the ratio $\sigma_1/\sigma_0$, which appears to be close to 1. If the position of the excited level, $\epsilon_m$, is fixed, the lineshape is asymmetric. When $\rho$ is close to 1, the PL lineshape also shows two maxima, however, there is a peculiar interplay in their magnitudes when the excitation energy is changed. This also matches the experimental data quite well. Such interplay can be easily understood. First (second) PL maximum corresponds to the distribution of ground-state energies of such subset of quantum dots, for which the lowest (highest) split-off level coincides with $E_{laser}$. When $E_{laser}$ is below $\overline E_m$, the amounts of first type of dots is larger than that of the second type. When $E_{laser}$ is larger than $\overline E_m$, the dots of the second type win. The interplay is absent in PLE, because for each fixed ground-state energy there is always equal amount of lowest and highest split-off levels. The magnitudes of $\sigma_1$ and $\sigma_0$ can be obtained independently as follows. For $\rho=0.94$, the maximum of $D_\rho(\epsilon;\sigma)$ lies at $\epsilon=0.765\sigma$. Then, from the position of the PLE maximum, $\sigma_1$=14.5 meV/0.765=19 meV. The spread of the ground states $\sigma_0$ can be determined independently from the nonselective PL [curve 9 in Fig.\ 6(a)]. It gives $\sigma_0$=18.2 meV. If we know the shape of the wave function, we may find the values for the parameters $\sigma_{0,1}$ and $\rho$, using Eqs.\ (2,4). As a guess, we may try the wave functions for the cylindrical quantum well with infinite walls: \begin{equation} \psi_m({\bf r})\propto\cos(\frac{\pi r_\bot}{h}) J_m(\frac{\nu_m r_\|}{R}), \end{equation} where $h$ and $R$ are correspondingly the thickness and the radius of the quantum dot, and $\nu_m$ is the root of the Bessel function $J_m$. The integrals of interest, $J_{mn}=\int d^3r\ \psi_m^2\psi_n^2$, are equal to: $J_{00}=2.098$, $J_{11}=1.552$, and $J_{01}=1.435$. It is easy to find all parameters in this approximation. First, let us estimate the magnitude of the spread. To find $\sigma_0$ one has to know the volume of the quantum dot. Taking it to be the volume of the cylinder with the thickness 2.5 nm and diameter 25 nm,\cite{P5} and using the parameters of InGaAs: $x_0=0.5$, lattice constant $a=0.585$ nm and $(dE_g/dx)=1.16$ eV (both at $x=0.5$),\cite{LB} we obtain $\sigma_0=13$ meV. This value is less than the experimental value of 18.2 meV. It, however, shows that at least a significant part of the spread is caused by the composition fluctuations. There are, of course, some other reasons for spreading. The total spread, however, cannot be less than the calculated value. Two remaining dimensionless parameters are: $\rho=0.795$, and $(\sigma_0/\sigma_1)=1.16$. Though the ratio $\sigma_0/\sigma_1$ is in a reasonable agreement with the experiment, the value of the correlation coefficient $\rho$ is significantly less than the experimental value of 0.94. Note that the dimensionless parameters depend only on the form of the wave functions. Though it is possible to relate the discrepancy in $\rho$ to the unknown shape of the real wave function, there is a more serious reason for this coefficient to be closer to 1. Among the other causes of spreading of energy levels, which are not taken into account in the presented theory, there are {\em fluctuations of the shape} of quantum dots. The distortion of the shape of a quantum dot, even when small, cannot be represented as a potential perturbation in the Schr\"odinger equation. It's effect on the positions and splitting of the energy level $E_m$ can, however, be described by an effective secular matrix $\delta H$ of the same form as Eq.\ (6). The only difference is that the matrix elements $u$ and $x+iy$ are not described by the Eqs.\ (7) any more. Instead, they are determined by the integrals of the derivative of the unperturbed wave function at the boundary of the quantum dot.\cite{ED} Thus, the effect of the shape fluctuations is only in renormalizing the parameters $\sigma_0$, $\sigma_m$, and $\rho$. It is important, however, that for the case of pure shape fluctuations, the parameter $\rho$, defined as a correlation coefficient between $u$ and $\epsilon_0$, is exactly equal to 1. The reason is that in an axially-symmetic quantum dot, the normal derivative of the wave function at the boundary is just a number rather than a function of coordinate. Therefore, there should be an exact relation between $u$ and $\epsilon_0$ {\em for each particular shape distortion}. It is natural to assume that the effective value of $\rho$ is closer to 1 when both mechanisms are involved, than it is in the case of pure composition fluctuations. \section{Conclusion} In the present paper a simple theory is developed, which allows to describe selective photoluminescence data from an array of quantum dots with random parameters. The theory explains large apparent Stokes-like shift between emission and absorption energies as the average distance between the ground and first excited energy levels in the dots. It is shown that existence of a random spread of the interlevel distances in the dots causes deviation of the positions of the maxima of the peaks seen in the PL and PLE spectra. Such deviation can make it difficult to determine the properties of the statistical distribution of energy levels in the array. It is suggested how the proper parameters of the statistical distribution may be obtained from the experimental data. The random shifts and splittings of energy levels caused by a white-noise random potential in the dots are studied. The density of states and mutual level-level distribution function are obtained for the case of axially symmetric quantum dot. The energy-level distribution and the resulting PLE and selectively-excited PL spectra appear to be close to that observed in the experiment (see Fig.\ 6). It is shown that the major part of the spread of energies observed in the experiment originates from the random potential caused by the composition fluctuations. It is also suggested that the random fluctuations of the shape of the dots also contribute to the spread. \section*{Acknowledgments} I am grateful to A. L. Efros for formulating the problem and for numerous illuminating discussions. I would like to thank P. M. Petroff for providing the experimental data prior to the publication. I appreciate important comment by M. E. Raikh and useful discussions with P. M. Petroff, E. I. Rashba, J. M. Worlock, and F. G. Pikus. This work was supported by the Center for Quantized Electronic Structures (QUEST) of UCSB under subagreement KK3017.
\section{Introduction} More than twenty years ago, Brezin, Le Guillou and Zinn-Justin (BGZ) studied the phase transition of a cubic anisotropic system by means of renormalization group equations \cite{BreGuZ1}. Within a $(4-\varepsilon)$-expansion, they found that to lowest nontrivial order in $\varepsilon$, the only stable fixed point for $N < N_c = 4$ is the ${\rm O}(N)$-symmetric one, where $N$ is the number of field components appearing in the cubic anisotropic model. They interpreted this as an indication that the anisotropy is irrelevant as long as $N$ is smaller than four. For $N > 4$, the isotropic fixed point destabilized and the trajectories crossed over to the cubic fixed point. Recently, our knowledge of perturbation coefficients of the renormalization group functions of the anisotropic system was extended up to the five-loop level by Kleinert and Schulte-Frohlinde \cite{KleSchu2}. Since the perturbation expansions are badly divergent, they do not directly yield improved estimates for the crossover value $N_c$ where the isotropic fixed point destabilizes in favor of the cubic one. An estimate using Pad\'{e} approximants \cite{KleSchu2} indicates $N_c$ to lie below $3$, thus permitting real crystals to exhibit critical exponents of the cubic universality class. For a simple $\phi^4$-theory, the Pad\'{e} approximation is known to be inaccurate. In fact, the most accurate renormalization group functions for that theory have been obtained by combining perturbation expansions with large-order estimates and using a resummation procedure based on Borel-transformations \cite{KleJan}--\cite{GuZin5}. It is the purpose of this paper to derive the large-order behavior of the renormalization group functions for the anisotropic $g {\phi}^4$-theory. In a forthcoming paper we will combine these results with the five-loop perturbation expansion of Kleinert and Schulte-Frohlinde to derive the precise value for the crossover value $N_c$. For the simple $g {\phi}^4$-theory, the large-order behavior of perturbation coefficients has been derived by Lipatov \cite{Lip1,Lip2}, BGZ \cite{BreGuZ2} and others \cite{Bog}--\cite{Par2} in a number of papers. The generalization to the ${\rm O}(N)$-symmetric case was given in \cite{BreGuZ2}. An equivalent method for calculating the large-order behavior is based on the observation that for a negative coupling constant Green functions possess an exponentially small imaginary part due to the fact that the ground state is unstable \cite{Lang,CollSop}. The imaginary part is associated with the tunneling decay rate of the ground state. It determines directly the large-order behavior of the perturbation coefficients via a dispersion relation in the complex coupling constant plane. In the semiclassical limit, the imaginary part of all Green functions can be calculated with the help of classical solutions called {\em instantons}. For a massless $g {\phi}^4$-theory in $d = 4$ space dimensions, these instantons can be found analytically. The imaginary part is a consequence of a negative frequency mode in the spectrum of the fluctuation operator, whose determinant enters the one-loop correction to the instanton contribution. McKane, Wallace and de Alcantara Bonfim \cite{WalBon} found a way to continue the results of the $g {\phi}^4$-theory in $d = 4$ to a field theory in $d = 4-\varepsilon$ dimensions. They proposed an extended dimensional regularization scheme for nonperturbative renormalizing the imaginary parts of vertex functions. In the present work we have to extend this scheme to the case of a ${\phi}^4$-theory with cubic anisotropy, where the energy functional has the following form: \begin{equation} \label{HN} H(\vec{\phi})=\displaystyle{\int}d^dx \left[\frac{1}{2} \sum_{i=1}^N {\phi}_i (-{\nabla}^2) {\phi}_i + \frac{u}{4} (\sum_{i=1}^N {\phi}_i^2)^2+ \frac{v}{4} \sum_{i=1}^N \phi_i^4 \right] \,\, . \end{equation} For $N = 2$, the corresponding model in quantum mechanics was first studied by Banks, Bender and Wu \cite{BaBeWu,BaBe} who used multidimensional WKB-techniques to derive the large-order behavior of the perturbation expansion for the ground state energy. In $1990$ Janke \cite{Jan} presented a more efficient calculation using a path integral approach. In the present work, this approach will be generalized to quantum field theory and extended by a careful discussion of the region near the isotropic limit $v \rightarrow 0$. This is important, since the infrared-stable cubic fixed point is expected to appear very close to the ${\rm O}(N)$-symmetric one. In fact, it will be sufficient to give the quantum-field theoretical generalization of \cite{Jan} in terms of an expansion about the isotropic case in powers of $v$. The paper is organized as follows. The method is developed by treating first the case $N=2$. In \mbox{Section $2$} we derive the Feynman rules for the power series expansion of all Green functions around the isotropic limit. In \mbox{Section $3$} we calculate the small-oscillation determinants for the transversal and longitudinal fluctuations. In \mbox{Section $4$} we use the extended renormalization scheme of \cite{WalBon} to find the full (real and imaginary) vertex functions, and derive the renormalization constants to one loop. In \mbox{Section $5$} we calculate the imaginary parts of the renormalization-group functions and thus the large-order behavior of the perturbation coefficients. In \mbox{Section $6$}, finally, we extend the results to the physically relevant case $N=3$. \section{Fluctuations around the isotropic instanton} For positive coupling constants $u$ and $v$, the system defined by the energy functional (\ref{HN}) is stable and the Green functions are real. On the other hand, if both coupling constants are negative the system is unstable and the Green functions acquire an imaginary part. The corresponding functional integrals can be calculated by an analytical continuation from positive to negative coupling constants, keeping the factor $\exp{[\int (-\frac{u}{4} (\sum_{i=1}^N \phi_i^2)^2 -\frac{v}{4} \sum_{i=1}^N \phi_i^4)\,d^4x]}$ real. We perform this analytical continuation by means of a joint rotation in the two complex planes, substituting $u \rightarrow u \exp{(i\theta)}$ and $v \rightarrow v \exp{(i\theta)}$, and rotating the azimuthal angle $\theta$ from $0$ to $\pi$. At the same time, we rotate the contour of integration in the field space. The convergence of the functional integrals is maintained by the field rotation $\phi \rightarrow \phi \exp{(-i{\theta}/4)}$. A natural parameter for the anisotropy of the system is the ratio $\delta = v /(u+v)$. The isotropic limit corresponds to $\delta = 0$. During the joint rotation of $u$ and $v$, the parameter $\delta$ remains constant. Thus, $\delta$ is a good parameter for the anisotropy at both positive and negative couplings $u$ and $v$. We shall use the coupling constants $g=u+v$ and $\delta$ for a calculation of the Green functions. These are given by the functional integrals \begin{eqnarray} \label{G2M} \lefteqn{G^{(2M)} (x_1,x_2,\ldots,x_{2M})_{i_1,i_2,\ldots,i_{2M}} = } \nonumber \\ \nonumber \\ & &\frac{\int D \phi \,\, \phi_{i_1} (x_1) \phi_{i_2} (x_2) \cdot \ldots \cdot \phi_{i_{2M}} (x_{2M}) \exp\{-H[\phi]\} }{\int D \phi \exp\{ -H[\phi]\} } \, , \nonumber \\ \end{eqnarray} where the subscripts $i_k$ run through the $N$ components of the field $\phi$. As explained in the introduction, we shall first study the case $N=2$ with the free energy functional \begin{eqnarray} \label{Hn2} \lefteqn{H[\phi_1,\phi_2] = } \nonumber \\ \nonumber \\ & &\!\!\!\!\!\!\!\! \int d^dx \left\{\frac{1}{2} \left[\phi_1 (-{\nabla}^2) \phi_1+ \phi_2 (-{\nabla}^2) \phi_2 \right]+\frac{g}{4} \left[\phi_1^4+ 2(1-\delta) \phi_1^2 \phi_2^2 + \phi_2^4 \right] \right\}. \end{eqnarray} When expressed in terms of the coupling constants $g$ and $\delta$, the Green functions possess an imaginary part for $g < 0$. For the reasons explained above, we shall derive an expansion of the imaginary parts of the Green functions (\ref{G2M}) around the isotropic case, i.~e., in powers of $\delta$: \begin{equation} \label{IMG} {\rm Im} \, G = \sum_{n=0}^{\infty} a_n \, {\delta}^n \left(\frac{A}{-g}\right)^{p(n)} \exp \left(\frac{A}{g}\right) [1+{\cal O}(g)]\, . \end{equation} Each power ${\delta}^n$ has its own $n$-dependent imaginary part. Given such an expansion, the large-order estimates for the coefficients of the powers $g^k$ follows from a dispersion relation in $g$: \begin{equation} \label{DR} G=\frac{1}{\pi}\int\limits_{-\infty}^0 d\bar{g}\, \frac{{\rm Im}\, G(\bar{g}+i0)}{\bar{g}-g}\, . \end{equation} For a general discussion of the relationship between imaginary parts and large-order behavior see, for example, ch.\ $17$ of \cite{book}. If the power-series expansion of $G$ is \begin{equation} \label{LO1} G=\sum_{k,n=0}^{\infty}G_{kn}\, g^k {\delta}^n \, , \end{equation} the coefficients $G_{kn}$ have the asymptotic behavior \begin{equation} \label{LO2} G_{kn}\stackrel{k \rightarrow \infty}{\longrightarrow}\, -\frac{a_n}{\pi}\left(-1\right)^k \left(\frac{1}{A}\right)^k k!\, k^{p(n)-1} \left[1+{\cal O} \left(1/k \right)\right]. \end{equation} These estimates apply to perturbation coefficients, in which the maximal power of $g$ is much larger than the power of $\delta$. Being interested in the region close to the isotropic limit, this restricted large-order estimation will be sufficient. An expansion of the Green function (\ref{G2M}) around the instanton yields exponentially small imaginary parts in both numerator and denominator. In order to isolate the imaginary part of the numerator, we simplify the denominator in (\ref{G2M}) and calculate first the imaginary part of the approximate Green functions \begin{eqnarray} \label{TilG} \lefteqn{\tilde{G}^{(2M)} (x_1,x_2,\ldots,x_{2M})_{i_1,i_2,\ldots,i_{2M}} = } \nonumber \\ \nonumber \\ \!\!\!\!\!& &\frac{\int D \phi \,\, \phi_{i_1} (x_1) \phi_{i_2} (x_2) \cdot \ldots \cdot \phi_{i_{2M}} (x_{2M}) \exp\{-H[\phi]\} }{\int D \phi \exp\{ -H_0 [\phi]\} }\, , \nonumber \\ \end{eqnarray} where the denominator contains only the free energy functional $H_0$. With the aim of calculating (\ref{IMG}), we expand the fields around the classical solution of the isotropic limit $\delta = 0$ for the space dimension $d = 4$. Thus we write: \begin{equation} \label{exp} \vec{\phi}=\vec{u}_L \phi_c+\vec{u}_L \xi+\vec{u}_T \eta= \left( \begin{array}{r} \cos{\vartheta} \\ \sin{\vartheta} \end{array} \right) \phi_c+ \left( \begin{array}{r} \cos{\vartheta} \\ \sin{\vartheta} \end{array} \right) \xi+ \left( \begin{array}{r} -\sin{\vartheta} \\ \cos{\vartheta} \end{array} \right) \eta \, , \end{equation} where $\phi_c$ is the well-known $g {\phi}^4$-instanton in four dimensions \begin{equation} \label{fcl} \phi_c=\left(\frac{8}{-g}\right)^{1/2} \frac{\lambda}{1+{\lambda}^2(x-x_0)^2} \, , \end{equation} and $\vartheta$ is the rotation angle of the isotropic instanton in the $(\phi_1,\phi_2)$-plane. The fields $\xi$ and $\eta$ correspond to the degrees of freedom orthogonal to the rotation of that instanton. The parameters $x_0$ and $\lambda$ are position and scale size of the instanton, respectively. Inserting the expansion (\ref{exp}) in $H(\phi_1,\phi_2)$, we obtain the expression: \begin{equation} \label{Hall} H(\phi_1,\phi_2)=H(\phi_{1c},\phi_{2c})+H_1+H_2+H_3+H_4 \, , \end{equation} where $H_i$ collects all terms in $\xi$ and $\eta$ of order $i$. They are given by: \\ (1) Linear terms: \begin{eqnarray} \label{H1} H_1(\xi,\eta) &=&-\frac{\varepsilon 4(2)^{1/2}}{(-g)^{1/2}} \int d^dx\frac{\lambda^3}{[1+\lambda^2(x-x_0)^2]^2} \xi \nonumber \\ & &+\frac{\delta}{(-g)^{1/2}}\, \sin^2(2\vartheta)\,8(2)^{1/2} \int d^dx \frac{\lambda^3}{[1+\lambda^2(x-x_0)^2]^3} \xi \nonumber \\ & &+\frac{\delta}{(-g)^{1/2}}\, \sin(4\vartheta) \,4(2)^{1/2} \int d^dx \frac{\lambda^3}{[1+\lambda^2(x-x_0)^2]^3} \eta \end{eqnarray} (2) Quadratic terms: \begin{eqnarray} \label{H2} H_2(\xi,\eta) &=&\frac{1}{2} \int d^dx \xi M_L \xi+ \frac{1}{2} \int d^dx \eta M_T \eta \nonumber \\ & &+\delta \, \sin^2(2\vartheta) \, 6 \int d^dx \frac{\lambda^2}{[1+\lambda^2(x-x_0)^2]^2}(\xi^2-\eta^2) \nonumber \\ & &+\delta \, \sin(4\vartheta)\, 6 \int d^dx \frac{\lambda^2}{[1+\lambda^2(x-x_0)^2]^2} \xi \eta \, , \end{eqnarray} \hspace*{3ex} where $M_L$ and $M_T$ are the operators \begin{equation} \label{MLT} M_L = -{\nabla}^2-\frac{24 \lambda^2}{[1+\lambda^2 (x-x_0)^2]^2}\, , \,\, M_T = -{\nabla}^2- \frac{ 8(1-\delta) \lambda^2}{[1+\lambda^2 (x-x_0)^2]^2} \end{equation} (3) Cubic terms: \begin{eqnarray} \label{H3} H_3(\xi,\eta) &=&-(-8g)^{1/2} \int d^dx \frac{\lambda}{1+\lambda^2(x-x_0)^2} \left[\xi^3+(1-\delta) \xi \eta^2 \right] \nonumber \\ & &-\delta \, \frac{\sin^2(2\vartheta)}{2} \, (-8g)^{1/2} \int d^dx \frac{\lambda}{1+\lambda^2(x-x_0)^2} (3 \eta^2 \xi-\xi^3) \nonumber \\ & &-\delta \, \frac{\sin(4\vartheta)}{4} \, (-8g)^{1/2} \int d^dx \frac{\lambda}{1+\lambda^2(x-x_0)^2} (\eta^3-3 \eta \xi^2) \end{eqnarray} (4) Quartic terms: \begin{eqnarray} \label{H4} H_4(\xi,\eta) &=& +\frac{g}{4}\int d^dx \left[\xi^4+ \eta^4+2(1-\delta) \xi^2 \eta^2 \right] \nonumber \\ & &-\delta \, g \, \frac{\sin^2(2\vartheta)}{8} \int d^dx (\xi^4+\eta^4-6\xi^2 \eta^2) \nonumber \\ & &+\delta \, g \, \frac{\sin(4\vartheta)}{4} \int d^dx (\xi \eta^3-\eta \xi^3) \end{eqnarray} The linear terms (\ref{H1}) contain factors $\varepsilon$ or $\delta$, since the expansion of the fields around the isotropic instanton is extremal only in the four-dimensional isotropic limit. The classical contribution of the instanton is \begin{eqnarray} \label{Hclas} \lefteqn{H(\phi_{1c},\phi_{2c}) = } \\ \nonumber \\ & &-\frac{\lambda^{\varepsilon}}{g} \frac{8\pi^2}{3} \left[1-\frac{\varepsilon}{2}\left(2+\ln \pi+\gamma \right) \right] -\frac{\lambda^{\varepsilon} \delta }{g} \, \frac{8\pi^2}{3} \frac{\sin^2(2\vartheta)}{2}+ {\cal O} \left(\frac{ \delta }{g}\, \varepsilon \right) \! . \nonumber \end{eqnarray} The first term in $H(\phi_{1c},\phi_{2c})$ is evaluated up to the first order in $\varepsilon$, because the one-loop renormalization will require replacing $1/g \rightarrow 1/g_{\rm r} + {\cal O}[f({\delta}_{\rm r})/{\varepsilon}]$, where $g_r$ is the renormalized version of the coupling constant $g$. A contribution of order $\varepsilon$ is not needed in the second, $\delta$-dependent term, where it would produce a further factor $\delta$, and thus be part of the neglected terms ${\cal O}(g)$ in (\ref{IMG}). Due to the anisotropy of the action, the fluctuation expansion (\ref{H1})--(\ref{Hclas}) contains $\vartheta$-dependent parts. However, all these terms are of order $\delta$ and can therefore be handled by straightforward perturbation theory. Note that the angle $\vartheta$ disappears when the isotropic instanton is directed along the coordinate axis. Then the isotropic instanton coincides with the exact solution in $d = 4$ for $\delta > 0$. This is also seen by inspecting the potential in (\ref{Hn2}). For $\delta > 0$, the term ${\phi}_1^4 + {\phi}_2^4$ is larger than $2 (1-\delta) {\phi}_1^2 {\phi}_2^2$, so that the ``tunneling-paths'' of extremal action are obviously straight lines along the coordinate axis. For the region $\delta < 0$, the exact instanton follows from the known fact that the action (\ref{Hn2}) is invariant under the orthogonal transformation: $${\phi}_1 =({\tilde{\phi_1}} + {\tilde{\phi_2}})/{\sqrt{2}} \, , \qquad {\phi}_2 =({\tilde{\phi_1}} - {\tilde{\phi_2}})/{\sqrt{2}} \, ,$$ with the new coupling constants: $$g=(2-{\tilde{\delta}}) {\tilde{g}}/2 \, , \qquad \delta = -\frac{2 \, \tilde{\delta}}{2-\tilde{\delta}} \, ,$$ satisfying $\delta < 0$ for $\tilde{\delta} > 0$. In contrast to the method in \cite{Jan}, the treatment of the fluctuations in a power series in $\delta$ does not allow us to deal with all modes perturbatively. Near the ${\rm O}(2)$-symmetric case, the Gaussian approximation for the rotation of the instanton becomes invalid, and the Gaussian integral must be replaced by an exact angle integration. The separation of the rotation mode must be done with the help of collective coordinates. The Jacobian of the relevant transformation can be deduced from the isotropic system. The field $\eta$ of small oscillations must not contain modes of $M_T$ with eigenvalues of the order $\delta$, since these would vanish for $\delta = 0$. The discussion of the longitudinal part is given in \cite{Wal}. In dimensional regularization, all eigenvalues of $M_L$ which would be zero for $d = 4$ are of order $\varepsilon$ in $4-\varepsilon$ dimensions. To avoid eigenvalues of the order $1/{\varepsilon}$ in the propagator for the longitudinal fluctuations, all collective coordinates of the four-dimensional case must be retained in $d$ dimensions. Therefore the field $\xi$ in (\ref{exp}) contains no modes of $M_L$ with eigenvalues proportional to $\varepsilon$. In order to calculate the fluctuation factor and to separate the almost-zero modes from $\det M_L$ and $\det M_T$, it is convenient to do a conformal transformation onto a sphere in $d+1$ dimensions leading to a discrete spectrum for the transformed differential operators $M_L$ and $M_T$ \cite{Drum,DrumSho}. Their products of eigenvalues can be given in terms of the Riemann $\zeta$-function which we easily expanded near $\varepsilon=0$. Simple $1/{\varepsilon}$-poles, which are characteristic of dimensional regularization, appear directly from the known singularity of that function. In place of the fields $\xi$ and $\eta$, we define the fields ${\Phi}_1 (\rho)$ and ${\Phi}_2 (\rho)$ on the unit sphere in $d+1$ dimensions: \begin{equation} \Phi_1={\sigma}^{1-d/2} \xi \, \quad \Phi_2={\sigma}^{1-d/2} \eta \, , \end{equation} where $\sigma = 2/(1+x^2)$. The instanton is supposed to be centered at the origin and rescaled by $\lambda$. The spherical operator corresponding to the differential operator ${\nabla}^2$ is \begin{equation} V_0=\frac{1}{2} L^2- \frac{1}{4} d(d-2)\, , \end{equation} where $L^2 =\sum_{a,b} ({\rho}_a {\partial}_b-{\rho}_b {\partial}_a)^2$ is the total angular momentum operator on the $(d+1)$-dimensional sphere. The eigenfunctions of $V_0$ are the spherical harmonics $Y_m^l(\rho)$ in $d+1$ dimensions \cite{Harm}. They satisfy \begin{equation} V_0\, Y^l_m (\rho)=-(l+\frac{1}{2} d-1)(l+\frac{1}{2} d)\, Y^l_m(\rho)\, , \end{equation} and have the degeneracy \begin{equation} \nu_l(d+1)=\frac{(2l+d-1) \Gamma(l+d-1)}{\Gamma(d) \Gamma(l+1)} \, . \end{equation} After the transformation, the angle-independent quadratic part of $H_2$ in (\ref{H2}) becomes \begin{equation} \frac{1}{2} \int d \Omega \Phi_1 (-V_0-6) \Phi_1 + \frac{1}{2} \int d \Omega \Phi_2 \left[-V_0-2 (1-\delta) \right] \Phi_2 \, , \end{equation} where $d{\Omega}$ is the surface element of the unit sphere in $d+1$ dimensions. After rewriting the entire energy functional(\ref{Hall}) in terms of the new fields, we can summarize the Feynman rules for the diagrammatic evaluation of the functional integrals as follows: \\ $(1)$ Propagators (from the $\vartheta$-independent part of $H_2$): \\ \\ \hspace*{2em}~longitudinal propagator: \begin{equation} \epsfxsize=2cm \epsfbox{diaa1.eps}\,\, =\, \left[\left(-V_0-6\right)^{'}\right]^{-1} \end{equation} \hspace*{2em}~transversal propagator: \begin{equation} \hspace*{6ex} \epsfxsize=2cm \epsfbox{diaa2.eps}\, \, =\, \left\{\left[-V_0-2(1-\delta) \right]^{'}\right\}^{-1} \end{equation} The prime indicates that the almost-zero modes are omitted when forming the inverse.\\ $(2)$ Tadpoles $(H_1)$: \begin{eqnarray} \epsfxsize=1.5cm \epsfbox{diab1.eps} &\raisebox{1ex}{=}&\!\! \raisebox{1ex}{$\displaystyle{ -\varepsilon \left(\frac{2}{-g}\right)^{1/2} \sigma^{-1+{\varepsilon}/2} \Phi_1}$} \nonumber \\ \nonumber \\ \epsfxsize=1.5cm \epsfbox{diab2.eps} &\raisebox{1ex}{=}&\!\! \raisebox{1ex}{$\displaystyle{ \left(\frac{2}{-g}\right)^{1/2} \delta \, \sin^2(2\vartheta) \sigma^{{\varepsilon}/2} \Phi_1}$} \nonumber \\ \nonumber \\ \epsfxsize=1.5cm \epsfbox{diab3.eps} &\raisebox{1ex}{=}&\!\! \raisebox{1ex}{$\displaystyle{ \left(\frac{1}{-2g}\right)^{1/2} \delta \, \sin(4\vartheta) \sigma^{{\varepsilon}/2} \Phi_2}$} \nonumber \\ \end{eqnarray} The bold dot stands for the $\vartheta$-dependence of the vertex. \\ $(3)$ Two-point vertices ($\vartheta$-dependent part of $H_2$): \\ \begin{eqnarray} \epsfxsize=2.5cm \epsfbox{diac1.eps}\,\, \raisebox{1ex}{+} \,\, \epsfxsize=2.5cm \epsfbox{diac2.eps} &\raisebox{1ex}{=}&\!\! \raisebox{1ex}{$\displaystyle{ \frac{3}{2}\sin^2(2\vartheta) \, \delta \, \left(\Phi_1^2-\Phi_2^2 \right)}$} \nonumber \\ \nonumber \\ \epsfxsize=2.5cm \epsfbox{diac3.eps} &\raisebox{1ex}{=}&\!\! \raisebox{1ex}{$\displaystyle{ \frac{3}{2} \sin(4\vartheta) \, \delta \, \Phi_1 \Phi_2}$} \nonumber \\ \end{eqnarray} $(4)$ Cubic vertices $(H_3)$: \begin{eqnarray} \epsfxsize=1cm \epsfbox{diad1.eps} &\raisebox{3ex}{=}& \!\! \raisebox{3ex}{$\displaystyle{ -(-2g)^{1/2}\sigma^{-{\varepsilon}/2} \Phi_1^3}$} \nonumber \\ \epsfxsize=1cm \epsfbox{diad2.eps} &\raisebox{3ex}{=}&\!\! \raisebox{3ex}{$\displaystyle{ -(-2g)^{1/2} (1-\delta)\sigma^{-{\varepsilon}/2} \Phi_1 \Phi_2^2}$} \nonumber \\ \epsfxsize=1cm \epsfbox{diad3.eps}\,\, \raisebox{3ex}{+} \,\, \epsfxsize=1cm \epsfbox{diad4.eps} &\raisebox{3ex}{=}& \!\! \raisebox{3ex}{$\displaystyle{ -(-2g)^{1/2} \, \delta \, \frac{\sin^2(2\vartheta)}{2} \sigma^{-{\varepsilon}/2}\left(3\Phi_2^2 \Phi_1-\Phi_1^3 \right)}$} \nonumber \\ \epsfxsize=1cm \epsfbox{diad5.eps}\,\, \raisebox{3ex}{+} \,\, \epsfxsize=1cm \epsfbox{diad6.eps} &\raisebox{3ex}{=}&\!\! \raisebox{3ex}{$\displaystyle{ -(-2g)^{1/2} \, \delta \, \frac{\sin(4\vartheta)}{4} \sigma^{-{\varepsilon}/2} \left(\Phi_2^3-3\Phi_2 \Phi_1^2 \right)}$} \nonumber \\ \end{eqnarray} $(5)$ Quartic vertices $(H_4)$: \begin{eqnarray} \epsfxsize=1cm \epsfbox{diae1.eps}\,\, \raisebox{2.5ex}{+} \,\, \epsfxsize=1cm \epsfbox{diae2.eps}\,\, \raisebox{2.5ex}{+} \,\, \epsfxsize=1cm \epsfbox{diae3.eps} &\raisebox{2.5ex}{=}& \!\! \raisebox{2.5ex}{$\displaystyle{ \frac{g}{4} \sigma^{-\varepsilon} \left(\Phi_1^4+\Phi_2^4+ 2(1-\delta) \Phi_1^2 \Phi_2^2 \right)}$} \nonumber \\ \nonumber \\ \epsfxsize=1cm \epsfbox{diae4.eps}\,\, \raisebox{2.5ex}{+} \,\, \epsfxsize=1cm \epsfbox{diae5.eps} &\raisebox{2.5ex}{=}& \!\! \raisebox{2.5ex}{$\displaystyle{ \frac{g}{4}\, \delta \,\sin(4\vartheta) \sigma^{-\varepsilon} \left(\Phi_1 \Phi_2^3- \Phi_2 \Phi_1^3 \right)}$} \nonumber \\ \nonumber \\ \epsfxsize=1cm \epsfbox{diae6.eps}\,\, \raisebox{2.5ex}{+} \,\, \epsfxsize=1cm \epsfbox{diae7.eps}\,\, \raisebox{2.5ex}{+} \,\, \epsfxsize=1cm \epsfbox{diae8.eps} &\raisebox{2.5ex}{=}&\!\! \raisebox{2.5ex}{$\displaystyle{ -\frac{g}{4} \, \delta \, \frac{\sin^2(2\vartheta)}{2} \sigma^{-\varepsilon} \left(\Phi_1^4 +\Phi_2^4-6\Phi_1^2 \Phi_2^2 \right)}$} \nonumber \\ \end{eqnarray} In order to obtain the leading corrections we must consider all connected diagrams of order ${\cal O}(1/g)$, ${\cal O}({\delta}/g)$, and ${\cal O}(1)$. The contributions of these diagrams are added to $H_c$ in (\ref{Hclas}). We consider first the contributions to ${\cal O}(1/g)$, which arise from the connected tree diagrams \begin{equation} \epsfxsize=2cm \epsfbox{diaf1.eps}\, , \qquad \epsfxsize=1.5cm \epsfbox{diaf2.eps}\, , \qquad \epsfxsize=1.5cm \epsfbox{diaf3.eps} \end{equation} and their ${\Phi}_1$ insertions of zeroth order in $g$: \begin{equation} \epsfxsize=1.5cm \epsfbox{diag1.eps}\, ,\quad \epsfxsize=2cm \epsfbox{diag2.eps}\, , \quad \epsfxsize=2.5cm \epsfbox{diag3.eps}\, \quad \ldots \end{equation} Since each ${\Phi}_1$-vertex produces an $\varepsilon$-factor, the smallest order in $\varepsilon$ is two. However, terms of order ${\varepsilon}^2/g$ are negligible for our calculation even after one-loop renormalization, since: $1/g \rightarrow 1/g_{\rm r} + {\cal O}[f({\delta}_{\rm r})/{\varepsilon}]$ and $\varepsilon \rightarrow 0$. The diagrams of ${\cal O}({\delta}/g)$ can be generated from those of ${\cal O}(1/g)$ by one of the substitutions: \begin{equation} \epsfysize=1cm \epsfbox{diah1.eps} \quad \raisebox{2ex} {$\Longrightarrow$} \quad \epsfysize=1cm \epsfbox{diah2.eps} \raisebox{2ex}{$\, \, .$} \end{equation} By inspection, we see that the only diagram with an $\varepsilon$-power smaller than two is given by \begin{equation} \epsfxsize=2cm \epsfbox{diai1.eps} \end{equation} which is of order $\varepsilon \, {\delta}/g$. After the renormalization, a term of order ${\delta}_{\rm r}$ appears. But this term can be neglected in comparison with the ${\delta}_{\rm r}/g_{\rm r}$-term from $H_c$ in expression (\ref{Hclas}). Hence all the tree diagrams would enter only in a higher-order calculation. Contributions to order $g^0$ can appear from one-loop diagrams. The only possible candidates are \begin{equation} \epsfxsize=7cm \epsfbox{diaj3.eps} \raisebox{3ex}{$\, \, ,$} \end{equation} where the $1/{\varepsilon}$-pole from the loop integration and the $\varepsilon$-factor of the ${\Phi}_1$-vertex cancel. Following the method of \cite{Wal}, we have found that these diagrams contribute to the coefficient of the imaginary part of $\tilde{G}^{(2M)}$ in (\ref{TilG}) a factor \begin{equation} c_1^L c_1^T \, =\exp(-5+\delta-{\delta}^2/2) \, . \end{equation} \section{Quadratic Fluctuation Determinants} The angle-independent quadratic form is decomposed into a longitudinal and a transverse part. For the longitudinal fluctuations we obtain the determinant \begin{equation} \label{Lprod} \left(\frac{\det V_L}{\det V_{0L}} \right)^{-\frac{1}{2}} \!\!= \left[\frac{\det(-V_0-6)}{\det(-V_0)} \right]^{-\frac{1}{2}} \!\!= \prod^{\infty}_{l=0} \left[\frac{(l+\frac{1}{2} d-3)(l+\frac{1}{2} d+2)}{ (l+\frac{1}{2} d-1)(l+\frac{1}{2} d)} \right]^{-\frac{1}{2} \nu_l(d+1)}_{,} \end{equation} which coincides with that appearing in the one-component $g{\phi}^4$ theory, and is therefore known \cite{Wal}. It contains a bound state at $l=0$, this being responsible for the expected imaginary part of $\tilde{G}^{(2M)}$, and $d+1$ almost-zero eigenmodes of order $\varepsilon$ associated with the dilatation and translation degrees of freedom of the instanton in $4-\varepsilon$ dimensions. Extracting these modes from the product (\ref{Lprod}) in the framework of collective coordinates \cite{ZittLan,GeSa}, we obtain the well-known formal replacement rule for the determinant \begin{equation} \left(\frac{\det V_L}{\det V_{0L}} \right)^{-\frac{1}{2}} \Longrightarrow J^{V_L} (2\pi)^{-\frac{(d+1)}{2}} c_2^L \end{equation} with \begin{equation} c_2^L=(2\pi)^{-1/2} 5^{5/2} \exp \left[ \frac{3}{\varepsilon}+\frac{3}{4}- \frac{7}{2} \gamma+\frac{3}{\pi^2} \zeta^{'}(2) \right] \, , \end{equation} and the Jacobian of the collective coordinates transformation: \begin{equation} J^{V_L}=\lambda^{(d-1)} \left(-\frac{16\pi^2 \lambda^{\varepsilon}}{ 15g} \right)^{\frac{d+1}{2}} [1+{\cal O}(\varepsilon,g)] \, . \end{equation} The expression for $c_2^L$ contains the Euler constant $\gamma$, and the derivative of the Riemann-zeta function $\zeta(x)$ at the value $x=2$. Note the simple pole in $\varepsilon$, which is related to the ultraviolet divergence on the one-loop level. The transverse fluctuations contain neither negative nor zero modes for $\delta$ larger than zero. The corresponding fluctuation determinant is given by \begin{eqnarray} \label{Tprod} \left(\frac{\det V_T}{\det V_{0T}} \right)^{-\frac{1}{2}} &=& \left\{\frac{\det \left[-V_0-2(1-\delta) \right]} {\det(-V_0)} \right\}^{-\frac{1}{2}} \nonumber \\ \nonumber \\ &=& \prod^{\infty}_{l=0} \left[\frac{(l+ \frac{1}{2} d-1)(l+\frac{1}{2} d)-2(1-\delta)} {(l+\frac{1}{2} d-1)(l+\frac{1}{2} d)} \right]^{-\frac{1}{2} \nu_l(d+1)}. \end{eqnarray} Just one zero eigenmode appears for $l=0$ in the isotropic limit $\delta \rightarrow 0$, due to the rotational invariance in that case. For $\delta > 0$, the numerator in (\ref{Tprod}) contributes a factor $1/{\sqrt{2 \delta}}$. To avoid this artificial zero-mode divergence for $\delta \rightarrow 0$ we separate, as announced earlier, the angular degree of freedom from the integral measure via the collective-coordinates method. Similar to the longitudinal case, we make the formal substitution \begin{equation} \left(\frac{\det V_T}{\det V_{0T}} \right)^{-\frac{1}{2}} \Longrightarrow J^{V_T} (2\pi)^{-\frac{1}{2}} c_2^T \end{equation} with \begin{equation} J^{V_T}=\left(-\frac{16\pi^2 \lambda^{\varepsilon}}{ 3g} \right)^{\frac{1}{2}} \, , \end{equation} which coincides with the Jacobian of the isotropic model. Since the integration interval for the angle $\vartheta$ is compact, no singularity appears in the limit $\delta \rightarrow 0$. It is useful to illustrate the appearance of a divergent factor $1/{\sqrt{2 \delta}}$ in a careless use of Gaussian integral. Assuming $\delta > 0$, we expand the angle-dependent classical action around $\vartheta = 0$ up to quadratic order. The ensuing Gaussian integral can be evaluated after replacing the finite integration interval $\vartheta \in [-\pi,\pi]$ by the noncompact one $\vartheta \in [-\infty,\infty]$. In this way, the integral \begin{displaymath} J^{V_T}(2\pi)^{-\frac{1}{2}} \int\limits_{-\pi}^{+\pi} d \vartheta \exp \left[-\lambda^{\varepsilon} \frac{8\pi^2}{3} \, \frac{\delta}{|g|} \, \frac{\sin^2(2\vartheta)}{2} \right] \end{displaymath} is evaluated to \begin{displaymath} \left(\frac{8\pi \lambda^{\varepsilon}}{3|g|} \right)^{\frac{1}{2}} \int\limits_{-\infty}^{+\infty} d \vartheta \exp \left[-\frac{1}{2} \left( \frac{16\pi^2 \lambda^{\varepsilon}}{3|g|} \, 2\delta \right) \vartheta^2 \right] =\frac{1}{\sqrt{2 \delta}} \end{displaymath} showing the spurious would-be zero-mode singularity, as the consequence of false rotation-mode treatment. Excluding the $l=0$~-mode in the numerator on the right-hand side of (\ref{Tprod}), we obtain for $c_2^T$: \begin{eqnarray} & \displaystyle{c_2^T}= & \!\!\! 2^{1/2} \exp \left\{-\frac{1}{2} \sum_{l=1}^{\infty} \frac{(2l+d-1) \Gamma(l+d-1)}{\Gamma(d) \Gamma(l+1)} \ln \left[ \frac{(l+\frac{d}{2}+1)(l+\frac{d}{2}-2)} {(l+\frac{d}{2}-1)(l+\frac{d}{2})} \right] \nonumber \right. \\ \nonumber \\ & & \left. -\frac{1}{2} \sum_{l=1}^{\infty} \frac{(2l+d-1) \Gamma(l+d-1)}{ \Gamma(d) \Gamma(l+1)} \ln \left[ 1+\frac{2\, \delta}{(l+\frac{d}{2}+1) (l+\frac{d}{2}-2)} \right] \right\}. \nonumber \\ \end{eqnarray} The first sum is the well-known contribution from the isotropic limit. The second sum is to be expanded in powers of $\delta$. For large $l$, the sum diverges as $\varepsilon$ tends to zero. In order to separate off the divergence as a simple $1/{\varepsilon}$-pole, we use the zeta-function regularization method described in \cite{Wal}. A somewhat tedious but straightforward calculation leads to \begin{equation} c_2^T=(2\pi)^{-1/6} 3^{1/2} \exp \Bigg[\frac{1}{3\varepsilon}(1-\delta)^2 +\frac{1}{4}-\frac{1}{2}\, \gamma + \frac{\zeta^{'} (2)}{\pi^2}-\frac{1}{9} \delta+ \frac{37}{81} {\delta}^2 +{\cal O} \left({\delta}^3 \right) \Bigg]\, . \end{equation} The series in front of the $\varepsilon$-singularity ends after the quadratic power of $\delta$, thereby leading to the exact $(1-\delta)^2/3$-coefficient of the simple $\varepsilon$-pole. Collecting all contributions to the one-loop expression of ${\rm Im} \, \tilde{G}^{(2M)}$, we obtain the imaginary parts \begin{eqnarray} \lefteqn{{\rm Im} \, \tilde{G}^{(2M)} (x_1,x_2, \ldots,x_{2M})_{i_1,i_2,\ldots,i_{2M}} = } \nonumber \\ \nonumber \\ \ & & \!\!\!\!\!\!\!\! -\int d \lambda \, d^dx_0 \left\{ c_1^L c_1^T c_2^L c_2^T (2\pi)^{-\frac{(d+1)}{2}} (2\pi)^{-1/2} J^{V_L} J^{V_T} \exp \left(\frac{\lambda^{\varepsilon}A}{g} \right) \nonumber \right. \\ \nonumber \\ & & \left. \mbox{} \!\!\!\!\!\!\!\! \times \prod^{2M}_{\nu=1} \phi_c(x_{\nu}) \,\, \frac{1}{2} \int\limits_0^{2\pi} d \vartheta u_{L\,i_1}(\vartheta) u_{L\,i_2}(\vartheta) \cdots u_{L\,i_{2M}} (\vartheta) \exp[a\sin^2(2\vartheta)] \right\} \end{eqnarray} with \begin{equation} \label{aA} a=\frac{\lambda^{\varepsilon} 8\pi^2}{3g} \, \frac{\delta}{2} \, , \quad A=\frac{8\pi^2}{3} \left[1-\frac{\varepsilon}{2}(2+\ln \pi+\gamma) \right] \end{equation} and $i_k=1,2$. The expression contains a standard factor $1/2$, since for symmetry reasons each saddle point contributes only one half of the Gaussian integral. It remains to perform the collective coordinates integration over the dilatation $(\lambda)$, translation $(x_0)$, and the rotation $(\vartheta)$ degree of freedom. Having obtained the imaginary parts of the Green functions, we go over to the bare imaginary part of the vertex functions. Taking the Fourier transform and excluding the $(2\pi)^d\, \delta(\sum_i q_i)$-factors, an amputation of the external legs of $\tilde{G}^{(2M)}(x_1,x_2,\ldots,x_{2M})_{i_1,i_2,\ldots,i_{2M}}$ leads to: \begin{eqnarray} \label{V2M} \lefteqn{{\rm Im} \,\Gamma_{\rm b}^{(2M)} (q_m)_{i_1,i_2,\ldots,i_{2M}} = } \nonumber \\ \nonumber \\ & &\!\!\!\!\! \displaystyle{-c_{\rm b}} \int\limits_0^{\infty} \frac{d\lambda}{\lambda} \left\{ \lambda^{d-M(d-2)} \left(-\frac{\lambda^{\varepsilon}8\pi^2}{3g} \right)^{(d+2+2M)/2} \!\!\! \exp \left(\frac{\lambda^{\varepsilon}A}{g} \right) \prod^{2M}_{\nu=1} \left(\frac{q_{\nu}}{\lambda}\right)^2 \! \tilde{\phi} \left( \frac{q_{\nu}}{\lambda} \right) \nonumber \right. \\ \nonumber \\ & & \left. \mbox{} \!\!\!\!\! \times \frac{1}{2} \int\limits_0^{2\pi} d \vartheta u_{L\,i_1}(\vartheta) u_{L\,i_2}(\vartheta) \cdots u_{L\,i_{2M}} (\vartheta) \exp[a\sin^2(2\vartheta)] \right\} [1+{\cal O}(\varepsilon,g)] \end{eqnarray} with \begin{equation} \tilde{\phi}(q)=2^{d/2} \pi^{\frac{(d-2)}{2}} 3^{1/2} \left|q\right|^{1-(d/2)} K_{\frac{d}{2}-1} (\left|q\right|) \, , \end{equation} where $K_\mu(z)$ is the modified Bessel-function. The coefficient $c_{\rm b}$ is given by \begin{eqnarray} &c_{\rm b}=2^{-2/3} 3^{1/2} \pi^{-11/3} \exp &\!\!\! \Bigg[ \frac{1}{3\varepsilon} \left(10-2\delta+{\delta}^2\right)+ \frac{4 \zeta^{'} (2)}{\pi^2}- 4\gamma -4+\frac{8}{9} \, \delta \nonumber \\ \nonumber \\ & &\mbox{} -\frac{7}{162} \, {\delta}^2 +{\cal O} \left( {\delta}^3 \right) \Bigg]. \end{eqnarray} An $\varepsilon$-singularity results from the ultraviolet divergence of one-loop diagrams. As discussed in the following section, this simple pole in $\varepsilon$ is removed by a conventional coupling constant renormalization. The contribution from the imaginary part of the denominator in equation (\ref{G2M}) follows for $M=0$ in (\ref{V2M}). We observe that it is of higher order in $g$ in comparison with the imaginary part of the numerator, and can therefore be neglected. So the imaginary part of (\ref{TilG}) supplies the desired imaginary part of (\ref{G2M}). \section{Renormalization} In the preceding section we have calculated the contribution of the quadratic fluctuations around the isotropic instanton. The almost-zero eigenvalues of translation, dilatation and rotation have been extracted and treated separately. Of course, the resulting expressions are useless, if we are not able to renormalize the theory. This is the most difficult additional problem which arises in the transition from quantum mechanics to higher-dimensional field theories. A systematic scheme to renormalize both the real and imaginary part of vertex functions for a $g{\phi}^4$ theory in $4-\varepsilon$ dimensions was introduced by McKane, Wallace and de Alcantara Bonfim. They calculated the full (real and imaginary) renormalization-group constants using an extended minimal subtraction scheme to one loop (the conventional one is given in \cite{HoVe,Hooft}). We have extended their method to the case of two coupling constants in view of applications to critical phenomena. For an illustration we consider first the four-point vertex function. The result (\ref{V2M}) of the one-loop calculation about the instanton reads: \begin{eqnarray} \label{pol} \lefteqn{{\rm Im} \, \Gamma_{\rm b}^{(4)}(q_m)_{i,j,k,l}= } \nonumber \\ \nonumber \\ & &\displaystyle{\sum_{n=0}^{\infty}} \left\{ \frac{(-{\delta}/2)^n}{n!} \frac{1}{2} \int\limits_0^{2\pi} d \vartheta u_{L\,i}(\vartheta) u_{L\,j}(\vartheta) u_{L\,k}(\vartheta) u_{L\,l}(\vartheta) [\sin(2\vartheta)]^{2n} \nonumber \right. \\ \nonumber \\ & &\left. \mbox{} \times \left[ -c_{\rm b} \int\limits_0^{\infty} \frac{d\lambda}{\lambda} \lambda^{\varepsilon} \left(-\frac{\lambda^{\varepsilon}A}{g}\right)^{\frac{d+6+2n}{2}} \!\!\! \exp \left(\frac{\lambda^{\varepsilon}A}{g} \right) \prod^{4}_{\nu=1} \left(\frac{q_{\nu}}{\lambda}\right)^2 \! \tilde{\phi} \left( \frac{q_{\nu}}{\lambda} \right) \right] \right\} \end{eqnarray} with $A=\frac{8{\pi}^2}{3} +{\cal O}(\varepsilon)$ and $i,j,k,l=1,2$. It remains to evaluate the integral over the parameter $\lambda$. This is a relict of the introduction of collective coordinates in the instanton calculation. The integral converges for small $\lambda$ due to the exponentially decreasing of modified Bessel-function $K_{d/2-1}$. For large $\lambda$, the product $(q_{\nu}/{\lambda})^2 \tilde{\phi}(q_{\nu}/{\lambda})$ behaves like \begin{eqnarray} \left(\frac{q_{\nu}}{\lambda}\right)^2 \! \tilde{\phi}\left(\frac{q_{\nu}}{\lambda}\right) \!\!&=&\!\!2^2 3^{1/2} \pi \Bigg\{ \left(\frac{q_{\nu}}{2\lambda}\right)^{\varepsilon} \left[1+\frac{1}{2}\varepsilon \left(\gamma-\ln \pi \right)+ {\cal O}\left({\varepsilon}^2\right) \right] \nonumber \\ & &+\left(\frac{q_{\nu}}{2\lambda}\right)^2 \Bigg[ \left(\frac{q_{\nu}}{2\lambda}\right)^{\varepsilon} \left(\frac{2}{\varepsilon}+\gamma-\ln \pi+{\cal O}(\varepsilon)\right) \nonumber \\ & &\quad -\left(\frac{2}{\varepsilon}+1-\gamma-\ln \pi +{\cal O}(\varepsilon) \right) \Bigg] +{\cal O}\left(\frac{q_{\nu}}{2\lambda}\right)^4 \Bigg\} \, , \end{eqnarray} so that the integral diverges logarithmically as $\varepsilon$ goes to zero. This divergence causes a $1/{\varepsilon}$-pole in the imaginary part. Since after the above approximation of $\tilde{\phi}\left(q_{\nu}/{\lambda}\right)$ the integral diverges for $\lambda \rightarrow 0$, a small $\lambda$ cutoff $\mu$ has to be introduced. In this way, the $\lambda$-integral in (\ref{pol}) takes the form \begin{equation} \label{scal} -c_{\rm b} \int\limits_{\mu}^{\infty} \frac{d\lambda}{\lambda} \lambda^{\varepsilon} \left(-\frac{\lambda^{\varepsilon}A}{g}\right)^{\frac{d+6+2n}{2}} \!\!\! \exp \left(\frac{\lambda^{\varepsilon}A}{g} \right) (2^2 3^{1/2} \pi)^4 \prod^{4}_{\nu=1} \left(\frac{q_{\nu}}{2\lambda} \right)^{\varepsilon} \left[1+{\cal O}\left(\frac{q_{\nu}}{2\lambda} \right) \right] \, . \end{equation} The coupling constant $g$ is chosen to lie on top of the tip of the left-hand cut in the complex $g$-plane. Being integrated only near $g \rightarrow 0^{-}$, the integral can be evaluated perturbatively in $g$. Using \begin{equation} \label{lambda} \int\limits_{\mu}^{\infty} \frac{d\lambda}{\lambda} \lambda^{a\varepsilon} \lambda^{\varepsilon} \exp\left[-\left(\frac{\lambda^{\varepsilon} A}{|g|}\right)\right]= \frac{1}{\varepsilon}\frac{|g|}{A} \mu^{a\varepsilon} \exp\left[-\left(\frac{\mu^{\varepsilon} A}{|g|}\right)\right] [1+{\cal O}(g)]\, , \end{equation} we get for (\ref{scal}) the expression \begin{equation} 2^8 3^2 \pi^4 \,c_{\rm b}\,\, \frac{1}{\varepsilon}\,\, \frac{g}{A}\, \left(-\frac{\mu^{\varepsilon}A}{g}\right)^{\frac{d+6+2n}{2}} \!\!\! \exp \left(\frac{\mu^{\varepsilon}A}{g} \right) \prod^{4}_{\nu=1} \left(\frac{q_{\nu}}{\mu} \right)^{\varepsilon}. \end{equation} After expanding the factor $(q_{\nu}/{\mu})^{\varepsilon}$ in powers of $\varepsilon$, we obtain the typical finite contribution $\ln(q_{\nu}/{\mu})$. Since this term can be omitted by a minimal subtraction, we are left with \begin{eqnarray} \label{gim} \lefteqn{{\rm Im} \, \Gamma_{\rm b}^{(4)}(q_m)_{i,j,k,l}= } \nonumber \\ \nonumber \\ & &2^8 3^2 \pi^4 \,c_{\rm b}\,\, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \left\{ \frac{(-2\, \delta)^n}{n!\Gamma(2n+3)} \left(\frac{g}{A}\right) \left(-\frac{\mu^{\varepsilon}A}{g}\right)^{\frac{d+6+2n}{2}} \!\!\! \exp \left(\frac{\mu^{\varepsilon}A}{g} \right) \nonumber \right. \\ \nonumber \\ & &\left. \mbox{} \times \left[ 3\Gamma\left(n+\frac{3}{2}\right)^2 \,\,(S_{ijkl}-\delta_{ijkl})+ \Gamma\left(n+\frac{1}{2}\right) \Gamma\left(n+\frac{5}{2}\right) \,\, \delta_{ijkl} \right] \right\}\!, \end{eqnarray} where \begin{displaymath} S_{ijkl}=\frac{1}{3} (\delta_{ij} \delta_{kl}+\delta_{ik} \delta_{jl}+ \delta_{il} \delta_{jk}) \end{displaymath} and \begin{displaymath} \delta_{ijkl} = \left\{ \begin{array}{l} 1\, , \quad i=j=k=l \\ 0\, , \quad {\rm otherwise}\, . \end{array} \right. \qquad (i,j,k,l=1,2) \end{displaymath} The tensor structure arises explicitly upon doing the integrals \begin{displaymath} \int\limits_0^{{\pi}/2} \sin^{2\alpha+1} (\vartheta) \cos^{2\beta+1} (\vartheta) \,d\vartheta = \frac{\Gamma(\alpha+1) \Gamma(\beta+1)}{2\,\, \Gamma(\alpha+\beta+2)}=\frac{1}{2} B(\alpha+1,\beta+1) \end{displaymath} for all combinations of indices in ${\rm Im} \,{\Gamma}_{ijkl}^{(4)}$. The first term in (\ref{gim}) contains only mixed index combinations, whereas the second term has only a contribution from equal indices. Now we can proceed to renormalize the bare results. The real part of ${\Gamma}_{ijkl}^{(4)}(q)$ is the perturbative one, and it is easily calculated to one loop. In terms of our coupling constants, the real parts ${\rm Re} \,{\Gamma}_{ijkl}^{(4)}(q)$ take the form \begin{eqnarray} \label{greal} \lefteqn{{\rm Re} \, \Gamma_{\rm b}^{(4)}(q_m)_{i,j,k,l}= } \nonumber \\ \nonumber \\ & &\!\!\!\!\! \left\{-6(1-\delta) g+\left[\frac{3}{4\pi^2\varepsilon} \left(10-14 \delta+4 {\delta}^2\right) +{\cal O}({\varepsilon}^0) \right] g^2 \mu^{-\varepsilon} \right\} \{ S_{ijkl}-\delta_{ijkl} \} \nonumber \\ \nonumber \\ & &\!\!\!\!\! +\left\{-6g+\left[\frac{3}{4\pi^2\varepsilon} \left(10-2 \delta+{\delta}^2\right) +{\cal O}({\varepsilon}^0) \right] g^2 \mu^{-\varepsilon} \right\} \, \delta_{ijkl}\, , \end{eqnarray} where $\mu$ is the arbitrary momentum scale introduced above in (\ref{scal}). In the absence of an instanton, the wave function renormalization has no one-loop contribution. In the consequence, the expression (\ref{greal}) is rendered finite by a coupling constant renormalization only. The subscript r of the coupling constants indicates renormalized expressions in the conventional perturbative manner, i.~e., \begin{eqnarray} g_{\rm r}\!\!\!&=&\!\!\!g\mu^{-\varepsilon}-\frac{1}{8\pi^2\varepsilon} \left(10-2\delta+{\delta}^2 \right) g^2 \mu^{-2\varepsilon}\, , \nonumber\\ \nonumber\\ \delta_{\rm r}\!\!&=&\!\!\delta+\frac{1}{8\pi^2\varepsilon} \left(-2\delta+{\delta}^2+{\delta}^3 \right) g \mu^{-\varepsilon} \nonumber \\ \end{eqnarray} Inserting this into (\ref{gim}) and (\ref{greal}) we find the perturbatively renormalized vertex functions \begin{equation} \label{gren} \Gamma_{\rm r}^{(4)}(q_m)_{i,j,k,l}=-6\mu^{\varepsilon} \left[F_1(\delta_{\rm r},g_{\rm r}) \left(S_{ijkl}-\delta_{ijkl}\right)+F_2(\delta_{\rm r},g_{\rm r})\delta_{ijkl} \right] \end{equation} with \begin{eqnarray} \lefteqn{F_1({\delta}_{\rm r},g_{\rm r})= } \nonumber \\ & & \!\!\!\!\!(1-{\delta}_{\rm r}) g_{\rm r}+i 2^7 3^2 \pi^4 \, c_{\rm r} \,\, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \frac{(-2{\delta}_{\rm r})^n}{n!} \frac{\Gamma\left(n+\frac{3}{2} \right)^2}{\Gamma(2n+3)} \left(-\frac{A}{g_{\rm r}}\right)^{\frac{d+4+2n}{2}} \!\!\! \exp \left(\frac{A}{g_{\rm r}} \right), \nonumber \\ \end{eqnarray} and \begin{eqnarray} &F_2&\!\!\!\!({\delta}_{\rm r},g_{\rm r})= \nonumber \\ &g_{\rm r}&\!\!\!\!+i 2^7 3 \pi^4 \, c_{\rm r} \, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \frac{(-2{\delta}_{\rm r})^n}{n!} \frac{\Gamma\left(n+\frac{1}{2} \right) \Gamma\left(n+\frac{5}{2} \right)} {\Gamma(2n+3)} \left(-\frac{A}{g_{\rm r}}\right)^{\frac{d+4+2n}{2}} \!\!\! \exp \left(\frac{A}{g_{\rm r}} \right). \nonumber \\ \end{eqnarray} The coefficient $c_{\rm r}$ is given by \begin{eqnarray} \label{cr} c_{\rm r}\!\!\! &=&\!\!\! c_{\rm b} \exp \left[ -\frac{A}{8 \pi^2\varepsilon} \left(10-2 {\delta}_{\rm r} +{\delta}_{\rm r}^2 \right) \right] \nonumber \\ \nonumber \\ &=&\!\!\! 3^{1/2} 2^{-2/3} \pi^{-2} \exp \bigg[\frac{1}{3\varepsilon} \left(10-2 {\delta}_{\rm r}+{\delta}_{\rm r}^2 \right)- \frac{1}{3\varepsilon} \left(10-2 {\delta}_{\rm r}+{\delta}_{\rm r}^2 \right) \nonumber \\ \nonumber \\ & & \qquad \qquad \qquad \quad +\frac{4\zeta^{'}(2)}{\pi^2}-\frac{7}{3}\gamma-\frac{2}{3} +{\cal O}({\delta}_{\rm r}) +{\cal O}({\delta}_{\rm r} \, g_{\rm r}) \bigg]. \end{eqnarray} The $1/{\varepsilon}$-pole terms cancel. Thus, $c_{\rm r}$ remains finite for $\varepsilon \rightarrow 0$. For the leading expansion (\ref{IMG}) we can take the perturbative renormalized expression (\ref{cr}) at the position ${\delta}_{\rm r} = 0$. The only remaining singularity is the $1/{\varepsilon}$-factor in ${\rm Im} \,{\Gamma}^{(4)}_{\rm r}(q)_{ijkl}$. It requires a further renormalization. In our special choice of coupling constants, the disastrous $k!$-divergence of perturbation series appears in the expansions coefficients of $g^k$. Since $\delta$ is a dimensionless anisotropy measure for positive as well as negative $g$, we have expanded the imaginary part of vertex functions around the isotropic case in a well defined manner. For the derivation of nonperturbative corrections to the renormalization-group functions and for later applications, however, it is convenient to avoid ratios of coupling constants. Therefore we proceed with $v_{\rm r} = g_{\rm r} \, {\delta}_{\rm r}$ and $g_{\rm r}$ instead of ${\delta}_{\rm r}$ and $g_{\rm r}$. Then $v_{\rm r}$ receive the rule of ${\delta}_{\rm r}$. Now we apply the extended minimal subtraction and perform a second (nonperturbative) renormalization step to eliminate the $1/{\varepsilon}$-pole term: \begin{eqnarray} \label{REN} g_{\rm R}\!\!&=&\!\!g_{\rm r}+i 2^7 3 \pi^4 c_{\rm r} \frac{1}{\varepsilon} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left( n+\frac{1}{2},n+\frac{5}{2} \right) \nonumber \\ & & \qquad \qquad \qquad \quad \times \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+4+4 n}{2}} \exp \left(\frac{A}{g_{\rm r}} \right) \Bigg]\, , \nonumber \\ \nonumber \\ \nonumber \\ v_{\rm R}\!\!&=&\!\!v_{\rm r}-i 2^7 3 \pi^4 c_{\rm r} \frac{1}{\varepsilon} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n \left(\frac{4n}{2n+3}\right) B \left(n+\frac{1}{2},n+\frac{5}{2} \right) \nonumber \\ & & \qquad \qquad \qquad \quad \times \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+4+4 n}{2}} \exp \left(\frac{A}{g_{\rm r}} \right) \Bigg]\, . \end{eqnarray} The subscript R denotes the fully renormalized coupling constants. The required wave function renormalization in nonperturbative terms give no leading contributions to the imaginary part (see below). Therefore the $1/{\varepsilon}$-singularity in (\ref{gren}) is removed by a nonperturbative coupling-constant renormalization only. We now proceed to investigate the two-step renormalization for the wave function. There are two additional complications for the ${\Gamma}^{(2)}$-functions. First, ${\rm Im} \,{\Gamma}^{(2)}$ is quadratically divergent. Second, there is a momentum-dependent divergence in the imaginary part of ${\Gamma}^{(2)}$. For the $g{\phi}^4$-theory it was shown in \cite{WalBon} that the undesirable momentum-dependence disappears during the nonperturbative renormalization process. This is comparable to the situation in conventional two-loop perturbation expansion. We continue to show that this non-obvious cancellation still works in the more complex case of two coupling constants and more than one field components. According to (\ref{V2M}) the imaginary part of the unrenormalized two-point vertex function reads \begin{eqnarray} {\rm Im} \, \Gamma_{\rm b}^{(2)}(q)_{i,j}\!\!&=&\!\! c_{\rm b} \int\limits_0^{\infty} \frac{d\lambda}{\lambda} \left\{ \lambda^2 \left(-\frac{\lambda^{\varepsilon}A}{g}\right)^{\frac{d+4}{2}} \!\!\! \exp \left(\frac{\lambda^{\varepsilon}A}{g} \right) \left[ \left(\frac{q}{\lambda}\right)^2 \! \tilde{\phi} \left( \frac{q}{\lambda} \right) \right]^2 \nonumber \right.\\ \nonumber \\ & &\left. \qquad \qquad \times \frac{1}{2} \int\limits_0^{2\pi} d \vartheta u_{L\,i}(\vartheta) u_{L\,j}(\vartheta) \exp\left[a \sin^2(2\vartheta)\right] \right\}. \end{eqnarray} The integration over the angle $\vartheta$ can be done: \begin{eqnarray} \lefteqn{\frac{1}{2} \int\limits_0^{2\pi} d\vartheta u_{L\,i}(\vartheta) u_{L\,j}(\vartheta) \exp\left[a \sin^2(2\vartheta)\right]} \nonumber \\ \nonumber \\ & &\!\!\!\!\!\!\! =\delta_{ij} \sum_{n=0}^{\infty} \frac{a^n}{n!}\, 4^n \frac{\Gamma\left(n+ \frac{1}{2} \right) \Gamma\left(n+\frac{3}{2} \right) }{\Gamma(2n+2)} =\delta_{ij} \sum_{n=0}^{\infty} \frac{a^n}{n!}\, 4^n B \left(n+ \frac{1}{2},n+\frac{3}{2} \right). \nonumber \\ \end{eqnarray} Thus, we get the usual form: ${\rm Im} \,{\Gamma}^{(2)}_{b\,\,ij} ={\delta}_{ij}\,{\rm Im} \,{\Gamma}^{(2)}_{\rm b}$. The wave function renormalization is contained in $\frac{\partial}{\partial q^2} {\Gamma}^{(2)}(q)$. Following a similar procedure as in the case of ${\Gamma}^{(4)}$, we obtain the imaginary part \begin{eqnarray} \label{WFim} \frac{\partial}{\partial q^2}\, {\rm Im} \,\Gamma_{\rm b}^{(2)}(q)\!\!&=&\!\! 2^4 \pi^2 3 \,c_{\rm b}\, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \left\{\frac{(-2\, \delta)^n}{n!} B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \nonumber \right. \\ \nonumber \\ & &\left. \mbox{} \times \left[\gamma-\frac{1}{2}+\ln \left(\frac{q}{2\mu}\right) \right] \left(-\frac{\mu^{\varepsilon}A}{g} \right)^{\frac{d+2+2n}{2}} \exp\left( \frac{\mu^{\varepsilon}A}{g}\right) \right\} \, . \nonumber \\ \end{eqnarray} Apart from the appearance of the $(1/{\varepsilon}) \ln(q/{\mu})$ term, a very similar expression is found for ${\rm Im} \,{\Gamma}^{(4)}_{\rm b}(q)$. In order to show the cancellation of this term during the second stage of renormalization we need the two-loop expression for the real part of $\frac{\partial}{\partial q^2} {\Gamma}^{(2)}(q)$ \begin{equation} \label{WFreal} \frac{\partial}{\partial q^2}\,{\rm Re}\,\Gamma_{\rm b}^{(2)}(q)= 1+q^{-2\varepsilon} \left(\frac{4}{3}-\frac{2}{3} \delta+\frac{1}{3} {\delta}^2 \right) g^2 \left[\frac{3}{4(8\pi^2)^2\varepsilon} +{\rm P}+{\cal O}(\varepsilon) \right] \, , \end{equation} where {\rm P} denotes some number which is the contribution of order ${\varepsilon}^0$. At $q={\mu}$, (\ref{WFreal}) is commonly defined as the inverse wave function renormalization constant $(Z^{\phi}_{\rm p})^{-1}$. In terms of the perturbative renormalized coupling constants $Z^{\phi}_{\rm p}$ is given by \begin{equation} \label{zp} Z_{\rm p}^{\phi}=1-\frac{1}{4(8\pi^2)^2\varepsilon} \left(4-2 {\delta}_{\rm r} +{\delta}_{\rm r}^2 \right) g_{\rm r}^2 \, , \end{equation} and together with (\ref{WFim}), we find after the first step of renormalization \begin{eqnarray} \frac{\partial}{\partial q^2}\,\Gamma_{\rm r}^{(2)}(q)\!\!&=&\!\! 1+\left(\frac{4}{3} g_{\rm r}^2-\frac{2}{3} v_{\rm r} g_{\rm r} +\frac{1}{3} v_{\rm r}^2 \right) \left[{\rm P}-\frac{3}{2(8\pi^2)^2} \ln \left(\frac{q}{\mu}\right) +{\cal O}(\varepsilon) \right] \nonumber \\ \nonumber \\ &+&\!\!i\,2^4 \pi^2 3 \,c_{\rm r}\, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \left\{\frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \nonumber \right. \\ \nonumber \\ & &\left. \mbox{} \times \left[\gamma-\frac{1}{2}+\ln \left(\frac{q}{2\mu}\right) \right] \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+2+4n}{2}} \exp\left( \frac{A}{g_{\rm r}}\right) \right\} \, . \end{eqnarray} We have again introduced the coupling constants $v_{\rm r}$ and $g_{\rm r}$ which are more convenient for practical applications. This expression reads in the fully renormalized couplings (\ref{REN}) \begin{eqnarray} \lefteqn{\frac{\partial}{\partial q^2}\,\Gamma_{\rm r}^{(2)}(q) = 1+\left(\frac{4}{3} g_{\rm R}^2-\frac{2}{3} v_{\rm R} g_{\rm R} + \frac{1}{3} v_{\rm R}^2 \right) \left({\rm P}-\frac{3}{2(8\pi^2)^2} \ln \left(\frac{q}{\mu}\right) \right)} \nonumber \\ \nonumber \\ & & \mbox{}+i\,2^4 \pi^2 3 \,c_{\rm r}\, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \Bigg\{\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n \nonumber \\ \nonumber \\ & & \qquad \times \Bigg[2^4 \pi^2 A \,\, B \left(n+\frac{1}{2},n+\frac{5}{2} \right)\, \frac{4(n+1)}{2n+3} \left({\rm P}-\frac{3}{2(8\pi^2)^2} \ln \left(\frac{q}{\mu}\right) \right) \nonumber \\ \nonumber \\ & & \qquad \quad +B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \left(\gamma-\frac{1}{2}-\ln2+\ln \left(\frac{q}{\mu}\right) \right) \Bigg] \nonumber \\ \nonumber \\ & & \times \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+2+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg\} \, . \end{eqnarray} By using the definition of the $B$-function in terms of Gamma-functions one can easily read off the cancellation of $(1/{\varepsilon}) \ln(q/{\mu})$ singularities for every power $n$ of the coupling constant $v_{\rm R}$. Therefore we are left with \begin{eqnarray} \frac{\partial}{\partial q^2}\,\Gamma_{\rm r}^{(2)}(q)\!\!&=&\!\! 1+\left(\frac{4}{3} g_{\rm R}^2-\frac{2}{3} v_{\rm R} g_{\rm R} +\frac{1}{3} v_{\rm R}^2 \right) \left[{\rm P}-\frac{3}{2(8\pi^2)^2} \ln \left(\frac{q}{\mu}\right) \right] \nonumber \\ \nonumber \\ &+&\!\! i\,2^4 \pi^2 3 \,c_{\rm r}\, \frac{1}{\varepsilon} \left(2^4 \pi^2 A\,\,{\rm P}+\gamma-\frac{1}{2}-\ln2 \right) \nonumber \\ \nonumber \\ &\displaystyle{\times}&\!\!\! \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+2+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg] \, . \nonumber \\ \end{eqnarray} Now we can define the nonperturbative minimally subtracted wave function renormalization to remove the $1/{\varepsilon}$-pole: \begin{eqnarray} \label{znp} Z_{{\rm np}}^{\phi}\!\!&=&\!\!1- i\,2^4 \pi^2 3 \,c_{\rm r}\, \frac{1}{\varepsilon} \left(2^4 \pi^2 A\,\,{\rm P}+\gamma-\frac{1}{2}-\ln2 \right) \nonumber \\ \nonumber \\ & &\qquad \times \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \nonumber \\ \nonumber \\ & &\qquad \qquad \quad \times \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+2+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg] \, . \end{eqnarray} The full renormalization constant is defined by $Z^{\phi}:=Z^{\phi}_{\rm p}\,Z^{\phi}_{{\rm np}}$ and follows from the product of (\ref{zp}) and (\ref{znp}). A further important Green function is associated with the composite field $(1/2) {\phi}^2$: \begin{equation} G^{(2,1)}(x_1,x_2;x_3)_{i,j}=\left\langle {\phi}_i(x_1) {\phi}_j(x_2) \frac{1}{2} {\phi}_k^2(x_3) \right\rangle \, . \end{equation} By a straightforward application of the techniques discussed in the preceding sections we obtain for the corresponding vertex function \begin{eqnarray} \lefteqn{ {\rm Im} \, \Gamma_{\rm b}^{(2,1)}(q_1,q_2;q_3)_{i,j}= -\frac{1}{2} c_{\rm b} \int\limits_0^{\infty} \frac{d\lambda}{\lambda} \Bigg\{ \left(-\frac{\lambda^{\varepsilon}A}{g}\right)^{\frac{d+6}{2}} \!\!\! \exp \left(\frac{\lambda^{\varepsilon}A}{g} \right)} \nonumber \\ \nonumber \\ & &\!\!\!\!\! \times \prod_{\nu=1}^2 \left[ \left(\frac{q_\nu}{\lambda}\right)^2 \! \tilde{\phi} \left( \frac{q_\nu}{\lambda} \right) \right] \hat{\phi} \left(\frac{q_3}{\lambda}\right) \frac{1}{2} \int\limits_0^{2\pi} d \vartheta u_{L\,i}(\vartheta) u_{L\,j}(\vartheta) \exp\left[a \sin^2(2\vartheta)\right] \Bigg\} \, , \nonumber \\ \end{eqnarray} where \begin{equation} \hat{\phi}(q)=2^{(d-2)/2}\pi^{(d-4)/2} 3 \left| q \right|^{(4-d)/2} K_{(d-4)/2}\left(\left|q \right|\right) \, . \end{equation} Now we can proceed as in the case of wave function renormalization. Performing the collective coordinates integration over $\lambda$ and $\vartheta$ we obtain \begin{eqnarray} {\rm Im} \,\Gamma_{\rm b}^{(2,1)}(q_1,q_2;q_3)_{i,j}\!\!&=&\!\! {\delta}_{ij} \, 2^4 \pi^2 3^2 \,c_{\rm b}\, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \left\{\frac{(-2\, \delta)^n}{n!} B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \nonumber \right. \\ \nonumber \\ & &\left. \!\!\! \times \left[\gamma+\ln \left(\frac{q_3}{2\mu}\right)\right] \left(-\frac{\mu^{\varepsilon}A}{g} \right)^{\frac{d+4+2n}{2}} \exp\left( \frac{\mu^{\varepsilon}A}{g}\right) \right\}, \nonumber \\ \end{eqnarray} which agrees with the usual form ${\Gamma}^{(2,1)}_{{\rm b}\, ij}= \delta_{ij}\, {\Gamma}^{(2,1)}_{{\rm b}}$. At $q_3=0$, $q_1^2=q_2^2={\mu}^2$, the vertex function ${\Gamma}^{(2,1)}(q_1,q_2;q_3)$ defines a third renormalization constant, called $Z^{\phi^2}$, via \begin{equation} {\Gamma}^{(2,1)} \equiv \left(Z^{\phi} Z^{\phi^2} \right)^{-1} \, . \end{equation} The perturbative $Z^{\phi^2}$ follows from the real part of ${\Gamma}^{(2,1)}$: \begin{equation} {\rm Re} \Gamma^{(2,1)}_{\rm b}=1+ g q_3^{-\varepsilon} \left(\frac{4}{3}-\frac{1}{3} \delta \right) \left(-\frac{3}{8\pi^2\varepsilon}+{\rm Q}+{\cal O}(\varepsilon) \right)\, , \end{equation} where ${\rm Q}$ is a number of order ${\varepsilon}^0$. In terms of the perturbative renormalized coupling constants, the one-loop expression of $Z_p^{\phi^2}$ is given by \begin{equation} \label{pPP} Z_{\rm p}^{\phi^2}=1+\frac{1}{8\pi^2\varepsilon} \left(4-\delta_{\rm r}\right) g_{\rm r} \, . \end{equation} Including ${\rm Re} \Gamma^{(2,1)}_{\rm b}$ and ${\rm Im} \Gamma^{(2,1)}_{\rm b}$ we obtain after the perturbative step of renormalization \begin{eqnarray} \Gamma_{\rm r}^{(2,1)}(q_1,q_2;q_3)\!\!&=&\!\! 1+\left(\frac{4}{3} g_{\rm r}-\frac{1}{3} v_{\rm r}\right) \left[{\rm Q}+\frac{3}{8\pi^2} \ln \left(\frac{q_3}{\mu}\right) +{\cal O}(\varepsilon) \right] \nonumber \\ \nonumber \\ &+&\!\!i\,2^4 \pi^2 3^2 \,c_{\rm r}\, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \left\{\frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \nonumber \right. \\ \nonumber \\ & &\left. \mbox{} \times \left[\gamma+\ln \left(\frac{q_3}{2\mu}\right) \right] \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+4+4n}{2}} \exp\left( \frac{A}{g_{\rm r}}\right) \right\} \, . \end{eqnarray} In the second step of renormalization the $(1/{\varepsilon})\ln (q/{\mu})$ singularities cancel, and we find \begin{eqnarray} &\Gamma&\!\!\!\!\!_{\rm r}^{(2,1)}(q_1,q_2;q_3)= 1+\left(\frac{4}{3} g_{\rm R}-\frac{1}{3} v_{\rm R} \right) \left[{\rm Q}+\frac{3}{8\pi^2} \ln \left(\frac{q_3}{\mu}\right) \right] \nonumber \\ \nonumber \\ & &\quad -i\,2^4 \pi^2 3^2 \,c_{\rm r}\, \frac{1}{\varepsilon} \left(\frac{8\pi^2}{3} {\rm Q}-\gamma+\ln2 \right) \nonumber \\ \nonumber \\ & &\quad \displaystyle{\times} \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+4+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg] \, . \nonumber \\ \end{eqnarray} Hence, the nonperturbative $Z^{\phi^2}$ is \begin{eqnarray} \label{npPP} \lefteqn{Z_{{\rm np}}^{\phi} Z_{{\rm np}}^{\phi^2}=1+ i\,2^4 \pi^2 3^2 \,c_{\rm r}\, \frac{1}{\varepsilon} \left(\frac{8\pi^2}{3} {\rm Q}-\gamma+\ln2 \right)} \nonumber \\ \nonumber \\ & &\times \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+4+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg]. \end{eqnarray} With respect to the leading expansion (\ref{IMG}) this is equal to $Z_{{\rm np}}^{\phi^2}$. \section{Renormalization group functions} After having calculated the imaginary parts of the renormalized couplings and of the renormalization constants $Z^{\phi}$ and $Z^{\phi^2}$, we are in the position to derive the nonperturbative renormalization group functions $\beta(g_{\rm R},v_{\rm R})$, $\gamma(g_{\rm R},v_{\rm R})$ and $\gamma_{\phi^2}(g_{\rm R},v_{\rm R})$ in the usual way. First we derive the $\beta$-functions from the expressions (\ref{REN}) for $g_{\rm R}(\mu)$ and $v_{\rm R}(\mu)$, respectively, and obtain: \begin{eqnarray} \lefteqn{\beta_g (g_{\rm R},v_{\rm R})=} \nonumber \\ & &\mu \frac{\partial}{\partial \mu} g_{\rm R} =\mu \frac{\partial}{\partial \mu} g_{\rm r}+ i 2^7 3 \pi^4 c_{\rm r} \frac{1}{\varepsilon} \displaystyle{\sum_{n=0}^{\infty}} \mu \frac{\partial}{\partial \mu} \Bigg[ \frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left( n+\frac{1}{2},n+\frac{5}{2} \right) \nonumber \\ \nonumber \\ & & \qquad \qquad \qquad \qquad \qquad \times \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+4+4 n}{2}} \exp \left(\frac{A}{g_{\rm r}} \right) \Bigg]\, , \nonumber \\ \lefteqn{\beta_v(g_{\rm R},v_{\rm R})=} \nonumber \\ & &\mu \frac{\partial}{\partial \mu} v_{\rm R} =\mu \frac{\partial}{\partial \mu} v_{\rm r}- i 2^7 3 \pi^4 c_{\rm r} \frac{1}{\varepsilon} \displaystyle{\sum_{n=0}^{\infty}} \mu \frac{\partial}{\partial \mu} \Bigg[ \frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{5}{2} \right) \nonumber \\ \nonumber \\ & & \qquad \qquad \qquad \qquad \qquad \times \left(\frac{4n}{2n+3}\right) \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+4+4 n}{2}} \exp \left(\frac{A}{g_{\rm r}} \right) \Bigg]\, . \end{eqnarray} The derivatives are taken at fixed bare couplings $g$ and $v$. These functions will be finite as $\varepsilon \rightarrow 0$. Furthermore, they will have a left-hand cut in the complex $g$-plane. On top of the tip of this cut we find \begin{eqnarray} \label{bet1} \beta_g (g_{\rm R},v_{\rm R})= \beta_g^{\rm p}(g_{\rm r},v_{\rm r}) \!\!\!\!&-&\!\!\!\!i 2^7 3 \pi^4 c_{\rm r} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left( n+\frac{1}{2},n+\frac{5}{2} \right) \nonumber \\ \nonumber \\ & & \times \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+6+4 n}{2}} \exp \left(\frac{A}{g_{\rm r}} \right) \Bigg] \, , \nonumber \\ \nonumber \\ \beta_v(g_{\rm R},v_{\rm R})= \beta_v^{\rm p}(g_{\rm r},v_{\rm r}) \!\!\!\!&+&\!\!\!\!i 2^7 3 \pi^4 c_{\rm r} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{5}{2} \right) \nonumber \\ \nonumber \\ & & \times \left(\frac{4n}{2n+3}\right) \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+6+4 n}{2}} \exp \left(\frac{A}{g_{\rm r}} \right) \Bigg] \, . \end{eqnarray} The real parts are the familiar perturbative $\beta$-functions \begin{eqnarray} \beta_g^{\rm p}(g_{\rm r},v_{\rm r})\!\!&=&\!\! -\varepsilon g_{\rm r}+\frac{1}{8\pi^2} \left(-2 g_{\rm r} v_{\rm r}+v_{\rm r}^2+10 g_{\rm r}^2 \right)\, , \nonumber \\ \beta_v^{\rm p}(g_{\rm r},v_{\rm r})\!\!&=&\!\! -\varepsilon v_{\rm r}+\frac{3}{8\pi^2} \left( 4g_{\rm r} v_{\rm r}-v_{\rm r}^2 \right)\, . \end{eqnarray} In the leading contributions the pair of coupling constants $(g_{\rm r},v_{\rm r})$ may be replaced by $(g_{\rm R},v_{\rm R})$ in (\ref{bet1}). This can be proven using (\ref{REN}). The result is \begin{eqnarray} \label{betaR} \beta_g (g_{\rm R},v_{\rm R})= \beta_g^{\rm p}(g_{\rm R},v_{\rm R}) \!\!\!\!&-&\!\!\!\!i 2^7 3 \pi^4 c_{\rm r} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left( n+\frac{1}{2},n+\frac{5}{2} \right) \nonumber \\ \nonumber \\ & & \times \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+6+4 n}{2}} \exp \left(\frac{A}{g_{\rm R}} \right) \Bigg]\, , \nonumber \\ \nonumber \\ \beta_v(g_{\rm R},v_{\rm R})= \beta_v^{\rm p}(g_{\rm R},v_{\rm R}) \!\!\!\!&+&\!\!\!\!i 2^7 3 \pi^4 c_{\rm r} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{5}{2} \right) \nonumber \\ \nonumber \\ & & \times \left(\frac{4n}{2n+3}\right) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+6+4 n}{2}} \exp \left(\frac{A}{g_{\rm R}} \right) \Bigg] \, . \end{eqnarray} Note that the $\beta$-functions have a form expected within a minimal subtracted scheme:~${\beta}_g(g_{\rm R},v_{\rm R})= -\varepsilon g_{\rm R} +{\beta}^{\, 4}_g(g_{\rm R},v_{\rm R})$ and ${\beta}_v(g_{\rm R},v_{\rm R})= -\varepsilon v_{\rm R}+{\beta}^{\, 4}_v(g_{\rm R},v_{\rm R})$, where the ${\beta}^{\, 4}$ denotes the $\beta$-functions in four dimensions. The critical large-distance behavior of the correlation functions is defined by the anomalous dimension of the field $\phi$, which is defined by \begin{equation} \gamma(g_{\rm R},v_{\rm R})=\mu \frac{\partial}{\partial \mu} \ln Z^{\phi}= \mu \frac{\partial}{\partial \mu} \ln Z_{\rm p}^{\phi}+ \mu \frac{\partial}{\partial \mu} \ln Z_{{\rm np}}^{\phi} \, . \end{equation} With (\ref{zp}) for $Z^{\phi}_{\rm p}$ and (\ref{znp}) for $Z^{\phi}_{{\rm np}}$ we get the $\varepsilon$-independent expression \begin{eqnarray} \label{gamR} \lefteqn{\gamma(g_{\rm R},v_{\rm R}) =} \nonumber \\ \nonumber \\ & &\!\!\!\!\!\frac{1}{2(8\pi^2)^2} \left(4 g_{\rm R}^2-2 v_{\rm R} g_{\rm R} +v_{\rm R}^2 \right) +i\,2^4 \pi^2 3 \,c_{\rm r}\, \left(2^7 \pi^4 3^{-1}\,{\rm P}+\gamma-\frac{1}{2}-\ln2 \right) \nonumber \\ \nonumber \\ & & \quad \times \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+4+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg] \, . \end{eqnarray} The divergence of the length scale at critical temperature is governed by the anomalous dimension of the composite field $\phi^2$: \begin{equation} \gamma_{\phi^2}(g_{\rm R},v_{\rm R})= -\mu \frac{\partial}{\partial \mu} \ln Z^{\phi^2}= -\mu \frac{\partial}{\partial \mu} \ln Z_{\rm p}^{\phi^2} -\mu \frac{\partial}{\partial \mu} \ln Z_{{\rm np}}^{\phi^2} \, . \end{equation} Using (\ref{pPP}) and (\ref{npPP}), we obtain \begin{eqnarray} \label{Npp2} \lefteqn{\gamma_{\phi^2}(g_{\rm R},v_{\rm R})=\frac{1}{8\pi^2} \left(4 g_{\rm R}-v_{\rm R} \right) +i\,2^4 \pi^2 3^2 \,c_{\rm r}\, \left(\frac{8\pi^2}{3} {\rm Q}-\gamma+\ln2 \right)} \nonumber \\ \nonumber \\ & & \quad \times \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n B \left(n+\frac{1}{2},n+\frac{3}{2} \right) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+6+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg]. \nonumber \\ \end{eqnarray} For the leading contributions to the imaginary part, it was possible to replace $(g_{\rm r},v_{\rm r})$ by $(g_{\rm R},v_{\rm R})$. We remark that the expansions (\ref{betaR}),(\ref{gamR}) and (\ref{Npp2}) have the correct isotropic limit for $v_{\rm R} \rightarrow 0$. \section{Generalization to the case $N=3$} Generalizing (\ref{exp}) to $N=3$ we write: \begin{eqnarray} \vec{\phi}\!\!\!&=&\!\!\!\vec{u}_L \phi_c+\vec{u}_L \xi+\vec{u}_{T_1} \eta+ \vec{u}_{T_2} \chi \nonumber \\ \nonumber \\ \!\!\!&=&\!\!\! \left( \begin{array}{c} \sin{\theta} \cos{\vartheta} \\ \sin{\theta} \sin{\vartheta} \\ \cos{\theta} \end{array} \right) (\phi_c+\xi)+ \left( \begin{array}{c} -\sin{\vartheta} \\ \cos{\vartheta} \\ 0 \end{array} \right) \eta+ \left( \begin{array}{c} \cos{\theta} \cos{\vartheta} \\ \cos{\theta} \sin{\vartheta} \\ -\sin{\theta} \end{array} \right) \chi \, , \end{eqnarray} where $\theta$ and $\vartheta$ are the rotation angles of the isotropic instanton. This change of variables yields the following expansion of the energy functional: \begin{eqnarray} \label{Hex3} H(\vec{\phi})=&H_c&\!\!\!+\frac{1}{2} \int d^dx \xi M_L \xi+ \frac{1}{2} \int d^dx \eta M_{T_1} \eta+ \frac{1}{2} \int d^dx \chi M_{T_2} \chi \nonumber \\ \nonumber \\ &-&\!\!\!\frac{\varepsilon 4(2)^{1/2}}{(-g)^{1/2}} \int d^dx\frac{\lambda^3}{[1+\lambda^2(x-x_0)^2]^2} \xi \nonumber \\ \nonumber \\ &-&\!\!\! (-8g)^{1/2} \int d^dx \frac{\lambda}{1+\lambda^2(x-x_0)^2} \left[\xi^3+(1-\delta)(\xi \eta^2+\xi \chi^2) \right] \nonumber \\ \nonumber \\ &+&\!\!\! \delta \,\, F(g,\vartheta,\theta,\xi,\eta,\chi) \end{eqnarray} with $M_L$ and $M_{T_1}=M_{T_2}=M_T$ of equation (\ref{MLT}). For brevity, we have written down explicitly only the terms responsible for the leading contributions in the expansion (\ref{IMG}). The remaining terms are denoted collectively by $\delta \,\, F(g,\vartheta,\theta,\xi,\eta,\chi)$. The classical contribution to the energy functional is \begin{eqnarray} H(\phi_{1c},\phi_{2c},\phi_{3c})\!\!\!&=&\!\!\! -\frac{\lambda^{\varepsilon}}{g} \frac{8\pi^2}{3} \left[1-\frac{\varepsilon}{2}\left(2+\ln \pi+\gamma \right) \right] \nonumber \\ \nonumber \\ & &\!\!\!\!\! -\frac{\lambda^{\varepsilon} \delta }{2g} \, \frac{8\pi^2}{3} \left[\sin^4{\theta} \sin^2(2\vartheta)+\sin^2(2\theta) \right]+ {\cal O}\left(\frac{ \delta }{g}\, \varepsilon \right). \end{eqnarray} With similar steps as for $N=2$ we obtain the bare imaginary parts of the vertex functions: \begin{eqnarray} \lefteqn{{\rm Im} \,\Gamma_{\rm b}^{(2M)} (q_m)_{i_1,i_2,\ldots,i_{2M}} = } \nonumber \\ \nonumber \\ & &\!\!\!\!\! \displaystyle{-c_{\rm b}^{(3)}} \int\limits_0^{\infty} \frac{d\lambda}{\lambda} \Bigg\{ \lambda^{d-M(d-2)} \left(-\frac{\lambda^{\varepsilon}8\pi^2}{3g} \right)^{(d+3+2M)/2} \!\!\! \exp \left(\frac{\lambda^{\varepsilon}A}{g} \right) \prod^{2M}_{\nu=1} \left(\frac{q_{\nu}}{\lambda}\right)^2 \! \tilde{\phi} \left( \frac{q_{\nu}}{\lambda} \right) \nonumber \\ \nonumber \\ & & \!\!\!\!\! \times \frac{1}{2} \int\limits_0^{2\pi} \int\limits_0^{\pi} \sin{\theta} d \theta d \vartheta \, \, u_{L\,i_1}(\vartheta,\theta) u_{L\,i_2}(\vartheta,\theta) \cdots u_{L\,i_{2M}} (\vartheta,\theta) \nonumber \\ \nonumber \\ & & \qquad \times \exp \left[a \sin^4{\theta} \sin^2(2\vartheta)+a \sin^2(2\theta) \right] \Bigg\} \left[1+{\cal O}(\varepsilon,g) \right] \, , \end{eqnarray} where \begin{eqnarray} c_{\rm b}^{(3)}\!\!\!&=&\!\!\! 2^{-5/6} 3 \pi^{-13/3} \exp \Bigg[ \frac{1}{3\varepsilon} \left(11-4\delta+2{\delta}^2\right)+ \frac{5 \zeta^{'} (2)}{\pi^2}- \frac{9}{2} \, \gamma -\frac{17}{4} \nonumber \\ \nonumber \\ & &\qquad \qquad +\frac{16}{9} \, \delta -\frac{7}{81} \, {\delta}^2 +{\cal O} \left( {\delta}^3 \right) \Bigg] \, , \quad (i_k\, =\, 1,2,3)\, , \end{eqnarray} with $a$, $A$ as in (\ref{aA}). After evaluating the integrals over the collective coordinates $\lambda$, $\theta$, and $\vartheta$ we obtain for the imaginary part of the four-point vertex function \begin{eqnarray} \lefteqn{{\rm Im} \, \Gamma_{\rm b}^{(4)}(q_m)_{i,j,k,l}= } \nonumber \\ \nonumber \\ & &2^8 3^2 \pi^4 \,c_{\rm b}^{(3)}\,\, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \Bigg\{ \frac{(-2\, \delta)^n}{\Gamma(2n+7/2)} \left(\frac{g}{A}\right) \left(-\frac{\mu^{\varepsilon}A}{g}\right)^{\frac{d+7+2n}{2}} \!\!\! \exp \left(\frac{\mu^{\varepsilon}A}{g} \right) \nonumber \\ \nonumber \\ & &\!\!\!\times \sum_{p=0}^n \frac{\Gamma(3+2n-p)\,\Gamma\left(p+\frac{1}{2} \right)} {(n-p)!\, p! \, \Gamma\left(2(n-p)+3\right)} \Bigg[ 3\Gamma\left(n-p+\frac{3}{2}\right)^2 \,\,(S_{ijkl}-\delta_{ijkl}) \nonumber \\ \nonumber \\ & & \qquad +\Gamma\left(n-p+\frac{1}{2}\right) \Gamma\left(n-p+\frac{5}{2}\right) \,\, \delta_{ijkl} \Bigg] \Bigg\}\, . \end{eqnarray} Similarly the bare imaginary parts of $\frac{\partial}{\partial q^2}\Gamma^{(2)}(q)$ and $\Gamma^{(2,1)}(q_1,q_2;q_3)$ are found to be \begin{eqnarray} \lefteqn{\frac{\partial}{\partial q^2} {\rm Im} \,\Gamma_{\rm b}^{(2)}(q)_{i,j}= \delta_{ij} \, 2^4 \pi^2 3 \,c_{\rm b}^{(3)} \left[\gamma-\frac{1}{2}+\ln \left(\frac{q}{2\mu}\right) \right]} \nonumber \\ \nonumber \\ & & \times \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \Bigg[\frac{(-2\, \delta)^n}{n!} \, I(n) \, \left(-\frac{\mu^{\varepsilon}A}{g} \right)^{\frac{d+3+2n}{2}} \exp\left( \frac{\mu^{\varepsilon}A}{g}\right) \Bigg] \, , \end{eqnarray} and \begin{eqnarray} {\rm Im} \,\Gamma_{\rm b}^{(2,1)}(q_1,q_2;q_3)_{i,j}\!\!&=&\!\! \delta_{ij} \, 2^4 \pi^2 3^2 \,c_{\rm b}^{(3)}\, \frac{1}{\varepsilon} \sum_{n=0}^{\infty} \left\{ \frac{(-2\, \delta)^n}{n!} I(n) \nonumber \right. \\ \nonumber \\ & &\left. \!\!\!\!\!\!\!\! \times \left[\gamma+\ln \left(\frac{q_3}{2\mu}\right)\right] \left(-\frac{\mu^{\varepsilon}A}{g} \right)^{\frac{d+5+2n}{2}} \exp\left( \frac{\mu^{\varepsilon}A}{g}\right) \right\}, \nonumber \\ \end{eqnarray} with \begin{equation} I(n) \equiv \sum_{p=0}^n \left( \begin{array}{c} n \\ p \end{array} \right) B \left(2+2n-p,p+\frac{1}{2}\right) B \left(n-p+\frac{1}{2},n-p+\frac{3}{2} \right). \end{equation} Now we apply the extended renormalization scheme. Using the coupling constants $v$ and $g$, we obtain for the nonperturbative renormalized couplings \begin{eqnarray} \label{c3R} g_{\rm R}\!\!\!&=&\!\!\!g_{\rm r} +i 2^7 3 \pi^4 c_{\rm r}^{(3)} \frac{1}{\varepsilon} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n J_g(n) \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+5+4 n}{2}} \exp \left(\frac{A}{g_{\rm r}} \right) \Bigg] \, , \nonumber \\ \nonumber \\ v_{\rm R}\!\!\!&=&\!\!\!v_{\rm r} -i 2^7 3 \pi^4 c_{\rm r}^{(3)} \frac{1}{\varepsilon} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm r}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n J_v(n) \left(-\frac{A}{g_{\rm r}} \right)^{\frac{d+5+4 n}{2}} \exp \left(\frac{A}{g_{\rm r}} \right) \Bigg]\, , \nonumber \\ \end{eqnarray} with \begin{eqnarray} \label{J} \displaystyle{ J_g(n)} \!\!\!&\equiv&\!\!\! \sum_{p=0}^n \left( \begin{array}{c} n \\ p \end{array} \right) B \left(3+2n-p,p+\frac{1}{2}\right) B \left(n-p+\frac{1}{2},n-p+\frac{5}{2} \right)\, , \nonumber \\ \nonumber \\ \displaystyle{ J_v(n)} \!\!\!&\equiv&\!\!\! \sum_{p=0}^n \Bigg\{ \left( \begin{array}{c} n \\ p \end{array} \right) \, \left[ \frac{4(n-p)}{2(n-p)+3} \right] \, B \left(3+2n-p,p+\frac{1}{2}\right) \nonumber \\ \nonumber \\ & & \qquad \qquad \qquad \qquad \quad \times B \left(n-p+\frac{1}{2},n-p+\frac{5}{2} \right) \Bigg\} \, , \end{eqnarray} and \begin{equation} \label{cc33} c_{\rm r}^{(3)}=3\,\, 2^{-5/6} \pi^{-5/2} \exp \bigg[ \frac{5\zeta^{'}(2)}{\pi^2}-\frac{8}{3} \, \gamma-\frac{7}{12} +{\cal O}({\delta}_{\rm r})+{\cal O}({\delta}_{\rm r} \, g_{\rm r}) \bigg]\, . \end{equation} For the nonperturbative renormalization constants $Z^{\phi}$ and $Z^{\phi^2}$ we find \begin{eqnarray} Z_{{\rm np}}^{\phi}\!\!&=&\!\!1- i\,2^4 \pi^2 3 \,c_{\rm r}^{(3)}\, \frac{1}{\varepsilon} \left(2^4 \pi^2 A\,\,{\rm P}+\gamma-\frac{1}{2}-\ln2 \right) \nonumber \\ \nonumber \\ & & \mbox{} \times \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n \, I(n) \, \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+3+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg] \, \end{eqnarray} and \begin{eqnarray} \lefteqn{Z_{{\rm np}}^{\phi} Z_{{\rm np}}^{\phi^2}=1+ i\,2^4 \pi^2 3^2 \,c_{\rm r}^{(3)}\, \frac{1}{\varepsilon} \left(\frac{8\pi^2}{3} {\rm Q}-\gamma+\ln2 \right)} \nonumber \\ \nonumber \\ & &\times \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n I(n) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+5+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg]. \end{eqnarray} To calculate the nonperturbative renormalization group functions $\beta(g_{\rm R},v_{\rm R})$, $\gamma(g_{\rm R},v_{\rm R})$ and $\gamma_{\phi^2}(g_{\rm R},v_{\rm R})$, we simply repeat the calculations of the case $N=2$. Using the expressions (\ref{c3R}) for $g_{\rm R}$ and $v_{\rm R}$ yields the $\beta$-functions \begin{eqnarray} \label{B3} \lefteqn{\beta_g (g_{\rm R},v_{\rm R})=} \nonumber \\ & &\beta_g^{\rm p}(g_{\rm R},v_{\rm R}) -i 2^7 3 \pi^4 c_{\rm r}^{(3)} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n J_g(n) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+7+4 n}{2}} \exp \left(\frac{A}{g_{\rm R}} \right) \Bigg]\, , \nonumber \\ \nonumber \\ \lefteqn{\beta_v(g_{\rm R},v_{\rm R})=} \nonumber \\ & &\beta_v^{\rm p}(g_{\rm R},v_{\rm R}) +i 2^7 3 \pi^4 c_{\rm r}^{(3)} \displaystyle{\sum_{n=0}^{\infty}} \Bigg[ \frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n J_v(n) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+7+4 n}{2}} \exp \left(\frac{A}{g_{\rm R}} \right) \Bigg]\, . \nonumber \\ \nonumber \\ \end{eqnarray} The expressions for the nonperturbative anomalous dimension of the field $\phi$ and the composite field $\phi^2$ are given by \begin{eqnarray} \label{G3} \gamma(g_{\rm R},v_{\rm R})\!\!\!\!&=&\!\!\!\! \gamma^{\rm p}(g_{\rm R},v_{\rm R}) +i\,2^4 \pi^2 3 \,c_{\rm r}^{(3)}\, \left(2^7 \pi^4 3^{-1}\,{\rm P}+\gamma-\frac{1}{2}-\ln2 \right) \nonumber \\ \nonumber \\ & & \quad \times \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n I(n) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+5+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg] \, , \end{eqnarray} and \begin{eqnarray} \label{Npp3} \gamma_{\phi^2}(g_{\rm R},v_{\rm R})\!\!&=&\!\! \gamma_{\phi^2}^{\rm p}(g_{\rm R},v_{\rm R}) +i\,2^4 \pi^2 3^2 \,c_{\rm r}^{(3)}\, \left(\frac{8\pi^2}{3} {\rm Q}-\gamma+\ln2 \right) \nonumber \\ \nonumber \\ \!\!&\times&\!\! \sum_{n=0}^{\infty} \Bigg[\frac{v_{\rm R}^n}{n!} \left(\frac{3}{4\pi^2}\right)^n I(n) \left(-\frac{A}{g_{\rm R}} \right)^{\frac{d+7+4n}{2}} \exp\left( \frac{A}{g_{\rm R}}\right) \Bigg] \, , \end{eqnarray} respectively. \section{Discussion and Conclusions} Our main results are the real and imaginary parts (\ref{B3})--(\ref{Npp3}) of the renormalization-group functions $\beta$, $\gamma$ and $\gamma_{\phi^2}$ in terms of the nonperturbative renormalized couplings $g_{\rm R}$ and $v_{\rm R}$. Via the dispersion relation (\ref{DR}), we obtain the large-order behavior (\ref{LO2}). It is useful to check the expansion terms proportional to powers of $1/|g|$ accompanying the $v^n$ in the imaginary part of each renormalization-group function. If the power of $1/|g|$ is denoted by $p(n)$, we note that $p(n)$ is the same as for the corresponding vertex function before the integration over the dilatation parameter $\lambda$. This integration with the measure $d\ln \lambda$ reduces $p(n)$ by $1$ [see (\ref{lambda})]. However this effect is compensated by the differentiation with respect to the logarithm of the scale parameter $\mu$ in the definition of the renormalization group functions. The power $p(n)$ in the vertex functions can simply be explained: For a $2M$-point vertex function with $k$ $\phi^2$-insertions there is first a power $(1/|g|)^{(M+k)}$ from the fields. This is multiplied by a factor $(1/|g|)^{(d+(N-1)+1)/2}$ from the $d$ translational, $(N-1)$ rotational, and one dilatational would-be zero modes. The classical contribution to the energy functional yields a factor $(1/|g|)^2$ for each power of the anisotropic constant $v$. Hence, for the renormalization group function which follows at the one-loop level from the vertex function $\Gamma^{(2M,k)}$, the value of $p(n)$ is given by \begin{equation} p(n)=\frac{d+N+2(M+k)+4n}{2} \, . \end{equation} Inserting the imaginary parts into a dispersion relation (\ref{DR}), we can estimate the large-order coefficients of the corresponding series expansions \begin{equation} f(g,v)=\sum_{k,n=0}^{\infty} f_{kn}\, g^k \, v^n \, , \end{equation} where $n \ll k$. The function $f$ stands for ${\beta}_g$, ${\beta}_v$, $\gamma$ and $\gamma_{\phi^2}$, respectively. The rules how to go from the imaginary parts of $f$ to $f_{kn}$ are known from (\ref{LO2}). For instance, the large-order coefficients of the $\beta$-functions for $N=3$ are given by \begin{eqnarray} \label{concl} \beta_{g, \,kn}\!\!&=&\!\!c_{\rm r}^{(3)}2^7 3 \pi^3 \left(\frac{3}{4\pi^2}\right)^n \frac{ J_g(n) }{n!} \nonumber \\ \nonumber \\ & &\qquad \times (-1)^k \left(\frac{3}{8\pi^2}\right)^k k! \, k^{2n+(d+5)/2} \left[1+{\cal O}(1/k)\right] \, , \nonumber \\ \nonumber \\ \beta_{v, \,kn}\!\!&=&\!\!-c_{\rm r}^{(3)}2^7 3 \pi^3 \left(\frac{3}{4\pi^2}\right)^n \frac{ J_v(n) }{n!} \nonumber \\ \nonumber \\ & &\qquad \times (-1)^k \left(\frac{3}{8\pi^2}\right)^k k! \, k^{2n+(d+5)/2} \left[1+{\cal O}(1/k)\right] \, , \end{eqnarray} with $J$, $c_{\rm r}^{(3)}$ from (\ref{J}) and (\ref{cc33}), respectively. The leading large-$k$ behavior can be checked by means of a simple combinatorial analysis. We start with a series in terms of the standard couplings $u$ and $v$ [see (\ref{HN})]. For illustration, we consider the contribution of the pure $u$-powers. It is known from a theory with only one coupling constant, that the coefficients have the large-order behavior \begin{equation} \label{lao} f_l \, u^l \longrightarrow \gamma \, (-\alpha)^l \, \Gamma(l+b+1)\, u^l \qquad (l \gg 1) \, . \end{equation} Changing over to our couplings ($u = g-v$), the right-hand side of (\ref{lao}) contributes \begin{equation} \gamma \, (-1)^{l-k}\, (-\alpha)^l \, \Gamma(l+b+1) \, \left(\begin{array}{c} l \\ k \end{array} \right) \, g^k \, v^{l-k} \, , \end{equation} with $$\left(\begin{array}{c} l \\ k \end{array} \right)=\frac{\Gamma(l+1)}{k!\, (l-k)!} \, .$$ Replacing $l$ by $l=n+k$, the contribution of $g^k v^n$ has the form \begin{equation} f_{kn}\, g^k \, v^n \longrightarrow \gamma \, (\alpha)^n \, (-\alpha)^k \, \frac{\Gamma(k+n+b+1)\, \Gamma(k+n+1)}{k! \, n!} \, g^k \, v^n \, . \end{equation} Now, for large $k$, the $\Gamma$-functions can be approximated by \begin{equation} \Gamma(k+\delta+1) \longrightarrow k! \, k^{\delta} \, \left(1+{\cal O}(1/k) \right) \, . \end{equation} Thus in the region $n \ll k$ we obtain \begin{equation} f_{kn} \sim (-\alpha)^k \, k! \, k^{2n+b} \, , \end{equation} in agreement with (\ref{concl}). As stated in the introduction, the stable cubic fixed point is expected to lie in the vicinity of the isotropic fixed point. Since $v$ becomes very small in this region, reasonable results should be obtained by resumming the $g$-series accompanying each power $v^n$. This will be done in a forthcoming publication. \newpage
\section{Introduction} \label{sec:egbrintroduction} \setcounter{equation}{0} Quantum cosmology studies the relation between the observed universe and its boundary conditions in the hope that a natural {\it theory} of the boundary condition might emerge (see \cite{Halliwell} for an outstanding review of this enterprise.) Assessment of a particular theory requires an understanding of its implications for the present day. To that end, this paper elaborates on work of Gell-Mann and Hartle \cite{TSA} and Davies and Twamley \cite{DT} by examining the observable consequences for the diffuse extragalactic background radiation (EGBR) of one possible class of boundary conditions, those that are imposed time symmetrically at the beginning and end of a closed universe \cite{TSA}, and sketches some of the considerable difficulties in rendering this kind of model credible. Assuming such difficulties do not vitiate the consistency of time symmetric boundary conditions as a description of our universe, the principal conclusion is that these boundary conditions imply that the bath of diffuse optical radiation from extragalactic sources be at least twice that due only to the galaxies to our past, and possibly much more. In this sense, observations of the EGBR are observations of the final boundary condition. This conclusion will be seen to follow (section \ref{sec:limit}) because radiation from the present epoch can propagate largely unabsorbed until the universe begins to recollapse \linebreak (\cite{DT}, and section \ref{sec:opacity}), even if the lifetime of the universe is very great. By time symmetry, light correlated with the thermodynamically reversed galaxies of the recollapsing phase must exist at the present epoch. The {\it minimal} predicted ``excess" EGBR in a universe with time symmetric boundary conditions turns out to be consistent with present observations (section \ref{sec:observations}), but improved observations and modeling of galactic evolution will soon constrain this minimal prediction very tightly. In addition, many physical complications with the ansatz that time symmetric boundary conditions provide a reasonable and consistent description of the observed universe will become apparent. Thus this work may be viewed as outlining some reasons why even if very long-lived, our universe is probably not time symmetric. The plan of the paper is as follows. Section \ref{sec:motivations}\ discusses a model universe that will define the terms of the investigation. Section \ref{sec:tsbc}\ provides some perspective on doing physics with boundary conditions at two times with an eye toward section \ref{sec:difficulties}, where some aspects of the reasonableness of two time boundary conditions not immediately related to the extragalactic background radiation are discussed. Section \ref{sec:opacity}\ generalizes and confirms work of Davies and Twamley \cite{DT} in showing that for processes of practical interest, our future light cone (FLC) is transparent all the way to the recollapsing era over a wide range of frequencies, even if the universe is arbitrarily long-lived. Section \ref{sec:limit}\ explains why this fact implies a contribution to the optical extragalactic background radiation in a universe with time symmetric boundary conditions in excess of that expected without time symmetry. In the course of this explanation, some rather serious difficulties will emerge in the attempt to reconcile time symmetric boundary conditions, and a transparent future light cone, into a consistent model of the universe which resembles the one in which we live. Section \ref{sec:observations}\ compares the predictions of section \ref{sec:limit}\ for the optical EGBR to models of the extragalactic background light and observations of it. Section \ref{sec:summation}\ is reserved for summation and conclusions. \subsection{Motivations and A Model} \label{sec:motivations} The possibility that the universe may be time symmetric has been raised by a number of authors \cite{Gold,Wheeler1,Wheeler2,Hawking1,Zeh1,Zeh2,Zeh3,KZ}. Of course, what is meant is not {\it exact} time symmetry, in the sense that a long time from now there will be another Earth where everything happens backwards. Rather, the idea is that the various observed ``arrows of time" are directly correlated with the expansion of the universe, consequently reversing themselves during a recontracting phase if the universe is closed. Of central interest is the thermodynamic arrow of entropy increase, from which other time arrows, such as the psychological arrow of perceived time or the arrow defined by the retardation of radiation, are thought to flow \cite[are some reviews]{PTA,Zeh2,HPM}. However, the mere reversal of the universal expansion is insufficient to reverse the direction in which entropy increases \cite{Schulman1,PTA,Wheeler1,Wheeler2,Penrose1,Page1,Hawking2,HLL}. In order to construct a quantum physics for matter in a recollapsing universe in which the thermodynamic arrow naturally reverses itself, it appears necessary to employ something like the time neutral generalization of quantum mechanics \cite{Cocke,ABL,griffiths84,Unruh,TSA} in which boundary conditions are imposed near both the big bang {\it and} the big crunch. These boundary conditions take the form of ``initial" and ``final" density operators\footnote{I retain this terminology even though these operators may not be of trace class, for example in the familiar case where the final ``boundary condition" is merely the identity operator on an infinite dimensional Hilbert space.} which, when CPT-reverses of one another, define what is meant here by a time symmetric universe.\footnote{In other words, the effective decoherence functional for matter in a time symmetric universe is a canonical decoherence functional $d(h,h') = {\rm tr}[\rho_{\omega} h^{\dagger}\rho_{\alpha} h']/({\rm tr}\rho_{\alpha}\rho_{\omega})$, with $\rho_{\omega} = ({\cal CPT})^{-1}\rho_{\alpha} ({\cal CPT})$ \cite{TSA}.} In such a model the collection of quantum mechanical histories is time symmetric in the sense that each history in a decohering set ({\it i.e.\ }a set of histories in which relative probabilities may be consistently assigned) occurs with the same probability as its CPT-reverse \cite{TSA,Page2}. (As CP violation is small, there is for many purposes no difference between CPT- and T-symmetry with a T invariant Hamiltonian.) I do not describe these ideas in further detail because very little of the {\it formalism} of generalized quantum theory will be directly applied in this paper, but it is worth mentioning that in order for the resulting time symmetric quantum theory to have non-trivial predictions, the initial and final density operators must not represent pure states.\footnote{As the first in an occasional series of comments directed to those familiar with the ideas of generalized quantum theory, this is because if the initial and final density operators are pure, at most two histories can simultaneously be assigned probabilities \cite{TSA,dhgqm,craig95}, {\it i.e.\ }the maximum number of histories in any weakly decohering set is two! Complete information about the history of the universe must be encoded in the boundary conditions. Details of the formalism of generalized quantum theory may be found in, {\it e.g.,} \cite{LesH}. Generalized quantum mechanics with boundary conditions at two times is discussed more extensively in \cite{TSA} and \cite{dhgqm,craig95}.} Interesting theories therefore have boundary conditions which are quantum statistical ensembles. The interest in applying this class of quantum theories (namely, theories with CPT-related boundary conditions) in cosmology lies in the idea that the manifest arrow of time we observe is an emergent property of the universe, and not built directly into its structure by asymmetric dynamical laws or an asymmetric choice of boundary conditions. Dynamical laws are believed to be CPT-symmetric, so an asymmetric choice of boundary conditions is usually cited as the explanation for the existence of a definite arrow of time which flows in the same direction throughout the observable part of spacetime \cite{PTA,Zeh2,HPM}. However, it is worth investigating whether this assumption is {\it required} of us by making the alternate, apparently natural ansatz that the boundary conditions on a closed universe are (in a relevant sense) equivalent at the beginning and end of time, instead of the more usual assumption that the initial condition is somehow special and the final condition one of ``indifference," {\it i.e.\ }determined entirely by the past. Another point of view is that, as noted in \cite{TSA}, these alternative choices are in some sense opposite extremes. It is therefore of interest to determine whether they are distinguishable on experimental grounds, employing time symmetric boundary conditions as a laboratory for testing the sensitivity of physical predictions to the presence of a final boundary condition. For the benefit of those eager to proceed to the definite physical predictions of sections \ref{sec:opacity}\ and \ref{sec:egbr}, I now specify a model in which they might be expected to arise. (The cautiousness of this statement is explained in the sequel.) Sections \ref{sec:tsbc}, \ref{sec:amplifications}, and \ref{sec:difficulties}\ elaborate on the physics expected in a universe with two-time, time symmetric boundary conditions; here I merely summarize what is required from those sections for a complete statement of the assumptions. The model of the universe considered here consists in: \begin{itemize} \item A fixed closed, homogeneous and isotropic background spacetime, {\it viz.} a $k=+1$ Friedmann-Robertson-Walker (FRW) universe. The evolution of the scale factor is determined from Einstein's equations by the averaged matter content of the universe. (Inhomogeneities in the matter content can be treated as additional matter fields on this background.) \item Boundary conditions imposed on the matter content through a canonical decoherence functional $d_{\alpha\omega}$ \cite{dhgqm} with CPT-related density operators $\rho_{\alpha}, \rho_{\omega}$ that describe the state of matter at some small fiducial scale factor, near what would in the absence of quantum gravitational effects be the big bang and big crunch, but outside of the quantum gravity regime. The matter state described by one of these density operators reflects the presumed state of the early universe, namely matter fields in apparent thermal equilibrium at the temperature appropriate to the fiducial scale factor and the total amount of matter in the universe. Spatial fluctuations should be consistent with present day large scale structure, say being approximately scale invariant and leading to an amplitude of order $10^{-5}$ at decoupling in order to be consistent with recent COBE results \cite{COBE}. Possible further conditions on $\rho_{\alpha}, \rho_{\omega}$ are discussed at the end of this subsection. \end{itemize} The essence that a choice of boundary conditions with these apparently natural characterisics intends to capture is that of a universe in which the cosmological principle holds, which is smooth (and in apparent thermal equilibrium) whenever it is small, and which displays more or less familiar behaviour when larger. Most of the conclusions of this investigation really only rely on these general properties, but for the sake of definiteness a fairly specific model which has the right general physical characteristics is offered. However, as will be repeatedly emphasized in the sequel, in models with boundary conditions at two times, not only does the past have implications for the future, but the future has implications for the past. Therefore, in any attempt to model the {\it observed} universe with time symmetric boundary conditions, we need to make sure it is possible to integrate them into a self-consistent picture of the universe as we see it today. This means in particular that we must be prepared for the possibility that the early universe in a time symmetric universe may have properties different from those expected in a universe with an initial condition only. The model boundary conditions sketched above are not intended to be so restrictive as to rule out such differences, and hence must be taken with a grain of salt or two. The issue is then whether or not these properties are consistent with observation. Indeed, the prediction of an ``excess" optical EGBR (to be discussed in section \ref{sec:egbr}) in a universe with time symmetric boundary conditions is precisely of this character. (Of course, the most extreme possibility is that a universe burdened with these boundary conditions would look {\it nothing at all} like the universe in which we actually live.) Variations on this theme will recur frequently in the following sections. With these boundary conditions \cite{TSA}, the time neutral generalized quantum mechanics of Gell-Mann and Hartle defines an effective quantum theory for matter in the universe which may be imagined to arise from some other, more fundamental, quantum cosmological theory of the boundary conditions. (Presumably, the fundamental quantum cosmological decoherence functional incorporates the quantum mechanics of the gravitational field as well, a complication that is not addressed in this paper.) In this connection it is perhaps worth mentioning that it was once claimed \cite{Hawking1} that the no-boundary proposal for the initial condition\cite[\cite{Halliwell} is an excellent review]{HH} implied just this sort of effective theory in that it appeared to require that the universe be smooth whenever it was small. Thus a fundamental theory of an initial condition {\it only} apparently could be decribed by an effective theory with {\it two-time} boundary conditions. However, this claim has since been recanted \cite{Hawking2} due to a mathematical oversight. More generally, the no-boundary wave function does not appear to be a good candidate for a boundary condition imposed time symmetrically at both ends of a closed universe because it is a pure state \cite{TSA}, which as noted above yield quantum theories in which essentially all physical information is encoded in the boundary conditions alone. For contrast, see \cite{KZ}, in which it is asserted that the only sensible way to interpret the Wheeler-DeWitt equation necessitates a boundary condition requiring the universe to be smooth whenever it is small. In order to allow definite predictions to be made, the final key assumption of the model is that, for a suitable class of physically interesting coarse grainings ({\it e.g.\ }coarse grainings defining the domain of classical experience, or local quantum mechanics experiments), the probabilities for such coarse grained physical histories unfold near either boundary condition (relative to the total lifetime of the universe) in a fashion insensitive to the presence of the boundary condition at the other end, {\it i.e.\ }as if the other boundary condition were merely the identity. As discussed in section \ref{sec:tsbc}, simple stochastic models \cite{Cocke,TSA,Schulman2,Schulman3,Schulman4} suggest that this holds for processes for which the ``relaxation time" of the process to equilibrium is short compared to the total time between the imposition of the boundary conditions (see section \ref{sec:tsbc}), which expectation is rigorously supported in the case of Markov processes \cite{Schulman5}. Therefore, for the sake of brevity this predictive ansatz shall often be referred to as the ``Relaxation Time Hypothesis" (RTH). A more careful statement of the RTH would identify the classes of coarse grained histories (presumably at least those decribing short relaxation time processes at times sufficiently close to, for instance, the big bang) for which conditional probabilities (when defined) are, in a universe with boundary conditions $\rho_{\alpha}$, $\rho_{\omega}$, supposed to be close to those of a universe with $\rho_{\omega}=1$.\footnote{Some subtleties regarding equivalences between boundary conditions at one and two times are being concealed here, for which see \cite{craig95}.} It would also define ``close" and ``short relaxation time process" more precisely. (Thus, a rigorous statement of what is meant by the RTH requires a definite mathematical model. Because the concerns of this investigation encompass a variety of complicated physical processes, with the entire universe considered as a single physical system, I do not attempt that here. In specific cases the intuitive content of the ansatz ought to be clear enough.) In order to exploit the RTH to its fullest, I shall, when convenient, assume also that the universe is close to the critical density, so that its lifetime is very long. This plausibly realistic assumption maximizes the possibility that a model of the kind given above accurately describes our universe. With these assumptions, this model universe might be expected to closely resemble the universe as it is observed today {\it if} most familiar and important physical processes are examples of such ``short relaxation time processes.'' In particular, under the assumption that they are, predictions we expect of a single initial condition $\rho_{\alpha}$ only ({\it i.e.\ }$\rho_{\omega}= 1$) can (by the RTH) be assumed to hold near the initial condition in the model with CPT-related initial and final conditions. Such predictions should include those regarding inflation, relative particle abundances, and the formation of large (and small) scale structure. Because of the CPT-related boundary condition at the big crunch, a qualitatively similar state of affairs is then expected in the recollapsing era, but CPT reversed. As the thermodynamic arrow is caused fundamentally by gravitational collapse driving initially smooth matter away from equilibrium \cite[for example]{Penrose1,Zeh2}, the arrow of entropy increase near the big crunch will run in the opposite direction to that near the big bang.\footnote{Some discussion of the state of the universe when it is large, which somehow must interpolate between these opposed thermodynamic arrows, is provided in sections \ref{sec:tsbc}, \ref{sec:amplifications}, and \ref{sec:difficulties}.} Observers on planets in the recollapsing phase will find their situation indistinguishable from our own, with all time arrows aligned in the direction of increasing volume of the universe. It is this interesting (if unconventional) state of affairs which leads to the conclusion that observations of the EGBR can reveal the presence of a final boundary condition that is CPT-related to the initial, even if the lifetime of the universe is very great. How this comes about is the topic of section \ref{sec:egbr}. Before proceeding, some points made in the preceeding paragraph require qualification. (Readers whose only interest is the extragalactic background radiation should procede directly to section \ref{sec:opacity}. This discussion, and that of the next subsection, are positioned here for the sake of unity and perspective.) First, the ``relaxation times" for many important physical processes in a recollapsing universe {\it do not} appear to be short compared to the lifetime of the universe, even if that lifetime is arbitrarily long. The result described in section \ref{sec:opacity}, that light from the present epoch can propagate unabsorbed into the recollapsing phase, is an excellent example of this \cite{DT,TSA}. Some physical difficulties with the consistency of time symmetry to which this gives rise will become apparent in section \ref{sec:amplifications}. Further examples of such physical difficulties relating to issues such as gravitational collapse, the consequent emergence of a familiar thermodynamic arrow, and baryon decay in a universe with time symmetric boundary conditions are discussed briefly in section \ref{sec:difficulties}. The self-consistency of such models is thus in doubt, with these kinds of complications constituting arguments against the possibility that our universe possesses time symmetric boundary conditions. That is, it appears that time symmetry is {\it not} consistent with the central predictive ansatz that physics near either boundary condition is practically insensitive to the presence of the other, and it appears likely that a universe with time symmetric, low (matter) entropy boundary conditions would look {\it nothing like} the universe in which we actually live. Nevertheless, the strategy of sections \ref{sec:opacity}\ and \ref{sec:egbr}\ is to assume that the time symmetric picture is consistent with the gross characteristics of the observed universe and see what it predicts. Because of the prediction of a diffuse optical extragalactic background radiation that is testably different from that in a universe which is not time symmetric, we are provided with a two pronged attack on the hypothesis of time symmetric boundary conditions as a realistic description of our universe: the lack of a self-consistent picture of the observed universe, and observations of the EGBR. Second, for completeness it should be mentioned that in the context of the time neutral generalized quantum mechanics assumed here, there are further restrictions on the viability of time symmetric models as a realistic description of our universe which arise from the requirements of decoherence and the emergence of approximately classical behaviour.\footnote{It is not obvious that boundary conditions of the character noted above satisfy these restrictions, but if sufficient conditions obtain for the RTH to hold then it is at least plausible that they do. This is because such requirements might be expected to be more severe the more strongly correlated are the detailed states of the expanding and recollapsing eras. This should become clearer in section \ref{sec:tsbc}.} These topics are discussed by Gell-Mann and Hartle \cite{TSA}. Third, it is important to note that the conclusions of section \ref{sec:egbr}\ regarding the EGBR depend essentially on the assumed {\it global} homogeneity and isotropy, {\it i.e.\ }the ``cosmological principle." Thus, while inflation (apparently a generic consequence of quantum field theory in a small universe) can be taken to be a prediction of the assumed boundary conditions if they are imposed when the universe is sufficiently small, the popular point of view (somewhat suspect anyway) that inflation provides an {\it explanation} of homogeneity and isotropy inside the horizon is not a prediction of the effective theory of the universe considered here; it is an assumption. However, the other good things inflation does for us would still qualify as predictions. One further point requires mention. In the conventional picture of inflation, the matter in the universe is in a pure state (say, some vacuum state) when small. After inflation and reheating the matter fields appear to be, according to local coarse grainings, in thermal equilibrium, the states of different fields being highly entangled. Furthermore, the correlations required to infer that the complete state is actually pure have been inflated away. Nevertheless, the quantum state is of course still pure \cite{CC}. As noted previously, pure states are not viable candidates for the initial and final conditions. Therefore, in order to fit the conventional picture of inflation into an effective theory of the kind considered in this paper, the $\rho_{\alpha}$ and $\rho_{\omega}$ defined above must be a local description in the sense that they not contain the information required to know that the state from which they were inferred was actually pure, {\it i.e.\ }they must coarse-grain this information away. (The alternative is to employ an inflationary scenario in which the universe is not required to be in a pure state.) The unpleasantness of this restriction could be taken as an argument against the use of CPT-related boundary conditions in a fundamental quantum cosmological theory. Finally, for the cognoscenti of generalized quantum mechanics \cite[for example]{LesH}, there is a related observation that is even more interesting. Field theoretic Hamiltonians are CPT-invariant, so that if the CPT-related $\rho_{\alpha}$ and $\rho_{\omega}$ depend only on the Hamiltonian (as for instance an exactly thermal density operator does), $\rho_{\alpha}$ and $\rho_{\omega}$ are actually {\it equal}. (The same observation holds for merely T-related boundary conditions if the Hamiltonian is T invariant. However, CPT seems the more relevant symmetry if the Hamiltonian and boundary conditions of the considered effective theory of the universe arise from some more fundamental theory.) This is a potential difficulty for the sample model given above. A simple extension of a result of Gell-Mann and Hartle \cite[section 22.6.2]{TSA} to the case where the background spacetime is an expanding universe shows that the only dynamics allowed with identical initial and final boundary conditions is trivial: for alternatives allowed in sets of histories in which probabilities may be assigned at all, the time dependence of probabilities for alternative outcomes is essentially {\it independent of the Hamiltonian}, and is due only to the expansion of the universe.\footnote{To be more precise, consider the projections which appear in sets of histories that decohere for some $\rho_{\alpha}=\rho_{\omega}.$ For time independent Hamiltonians, time dependence of probabilities for such projections arises only if the projections are time-dependent in the Schr\"{o}dinger picture. Due to the expansion, this will be the case for many projections onto (otherwise time-independent) quantities of physical interest in an expanding universe.} In the present case this may be interpreted as a prediction that a universe which is required to be in thermal equilibrium whenever it is small {\it must remain so when large,} at least in the context of the theoretical structure of Gell-Mann and Hartle. (Classically the expectation is the same, so this result can hardly be written off as an artifact of the formalism.) Now, the evidence suggests that matter in the the early universe {\it was} in local thermal equilibrium. However, the inhomogeneities in the matter required to generate large scale structure must also be described by the boundary conditions, and, even in the usual case where there is only an initial condition, it is after all gravitational condensation which drives the appearance of a thermodynamic arrow.\footnote{More precisely, smoothly distributed matter in thermal equilibrium is {\it not} an equilibrium state when the gravitational field is included. Equilibrium states in the presence of gravitation have clumpy matter in them.} In the present case (the model with the two-time boundary conditions characterized above), the meaning of these observations is that one way to break the CPT invariance of the boundary conditions is for the specification of the deviations from perfect homogeneity and isotropy to be CPT non-invariant, in order that the theory admit interesting dynamics. As a specific example, I briefly consider the common circumstance where cosmological matter is given a hydrodynamic description. Scalar-type adiabatic pertubations may be completely specified by, {\it e.g.,} the values of the energy density perturbation and its (conformal-) time derivative over a surface of constant conformal time \cite{MFB}. (Scalar-type pertubations are the ones of interest, as they are the ones which may exhibit instability to collapse.) As it is intended here that gravity is treated classically, such a specification might come in the form of a probability distribution for approximately scale invariant metric and energy density fluctuations averaged over macroscopic scales. If this probability distribution reflects the underlying FRW homogeneity and isotropy (and is thus in particular P-invariant), in order to break T-invariance the distribution must distinguish between opposing signs of the (conformal-) time derivatives of the energy density perturbations. An alternative way to break the equality of CPT-related boundary conditions is for the locally observed matter-antimatter asymmetry to extend across the entire surface of constant universal time. (The recollapsing era will then be antimatter dominated in a CPT symmetric universe, consequently requiring that there be some mechanism to permit baryon decay.) This possibility may appear more attractive, but it is not without significant complications. These are addressed briefly at the end of section \ref{sec:amplifications}. \subsection{Physics with Time Symmetric Boundary Conditions} \label{sec:tsbc} As noted above, the time neutral generalized quantum mechanics of Gell-Mann and Hartle \cite{TSA} with CPT-related initial and final density operators yields a CPT symmetric ensemble of quantum mechanical histories, in the sense that each history in the collection is accompanied with equal probability by its CPT-reverse. The {\it collection} of histories (in each decohering set) is therefore statistically time symmetric (STS). Now, a set of histories can be statistically time symmetric without any individual history possessing qualitatively similar (if time-reversed) characteristics near both ends. Completely time asymmetric histories can constitute a statistically time symmetric set if each history in the collection is accompanied by its time reverse \cite{TSA,Page2}. With two-time boundary conditions, however, more is required. That is, there is a distinction between a CPT invariant {\it set} of histories, in which each history in a decohering set is accompanied, with equal probability, by its CPT reverse, and the considerably more restrictive notion of a statistically time symmetric universe studied here. In virtue of the boundary conditions at both the beginning and the end of time, probable histories (in a set which decoheres with these boundary conditions) will in general have both initial and final states which to some extent resemble the boundary conditions $\rho_{\alpha}, \rho_{\omega}$ \cite{craig95}. Perhaps the clearest way to understand this is to consider the construction of a classical statistical ensemble with boundary conditions at two times. The probability for each history in the ensemble may be found in the following way \cite{Cocke,TSA}. The boundary conditions are given in the form of phase space probability densities for the initial and final states of the system. Pick some initial state. Evolve it forward to the final time. Weight this history by the product of the probabilities that it meets the initial and final conditions, and divide by a normalizing constant so that all the probabilities sum to one. Thus, roughly speaking, in order for a history to be probable in this two-time ensemble both its initial and final state must be probable according to the initial and final probability densities, respectively. (For a deterministic classical system, two-time statistical boundary conditions are equivalent to a boundary condition at one time constructed in the obvious way. In quantum mechanics this is no longer true due to the non-commutability of operators at different times. Thus, while the described algorithm is merely a useful heuristic for understanding the implications of two-time boundary conditions {\it classically}, it is more essential quantum-mechanically.) What is not so clear is that the resulting probable histories look anything like the probable histories in the ensemble with just, say, the initial condition. In general they will not. Classically, this is merely the statement that the probability measure on a space of classical deterministic solutions defined by two-time boundary conditions will in general be quite different than the measure defined by the initial probability density only. For classical systems with stochastic dynamics or for quantum statistical systems, the simple models studied by Cocke \cite{Cocke}, Schulman \cite{Schulman2,Schulman3,Schulman4,Schulman5}, and others yield some insight into what is required for evolution near either boundary condition to be influenced only by that boundary condition. Here I merely summarize the intuitively transparent results of this work in the context of time symmetric boundary conditions. In particular, the emphasis will be on boundary conditions which represent ``low entropy" initial and final states. In the absence of a final condition, an initially low entropy state generally implies that evolution away from the initial state displays a ``thermodynamic arrow" of entropy increase (relative to the coarse graining defining the relevant notion of entropy; see \cite{Zeh2} for a pertinent overview.) In an ensemble with two-time, time symmetric, low entropy boundary conditions, under what conditions will a familiar ``thermodynamic arrow" appear near either boundary condition? A pertinent observation is that, roughly speaking, in equilibrium all arrows of time disappear. Therefore, if the total time between the imposition of the boundary conditions is much longer than the ``relaxation time" of the system to equilibrium (defined as the characteristic time a similar system with an initial condition only takes to relax), it might be expected that entropy will increase in the familiar fashion away from either boundary condition. That is, following the coarse-grained evolution of the system forward from the initial condition, entropy increases at the expected rate until the system achieves equilibrium. The system languishes in equilibrium for a time, and then entropy begins to decrease again until the entropy reaches the low value demanded by the final condition. Thus, this is a system in which the ``thermodynamic arrow of time" reverses itself, which reversal is enforced by the time symmetric, low entropy boundary conditions. There is no cause to worry about the coexistence of opposed thermodynamic arrows. Histories for which entropy increases away from either boundary condition are readily compatible with both boundary conditions, and the probable evolutions are those in which, {\it near either boundary condition}, there is a familiar thermodynamic arrow of entropy increase away from the nearest boundary condition. There, the other boundary condition is effectively invisible. Essentially this is because in equilibrium all states compatible with constraints are equally probable; the system ``forgets" its boundary conditions. On the other hand, if there is not time enough for equilibrium to be reached there must be some reconciliation between the differing arrows of increasing entropy \cite{Cocke,PTA,Penrose1,Penrose2,TSA,Schulman5}. As only histories which satisfy the required boundary conditions are allowed, it can be anticipated that the statistics of physical processes that would ordinarily (in the absence of the final boundary condition) lead to equilibration would be affected because histories which were probable with an initial condition only are no longer compatible with the final condition. The rate of entropy increase is slowed, and in fact the entropy may never achieve its maximum value. In other words, the approach to equilibrium is suppressed by the necessity of complying with the low entropy boundary condition at the other end. More probable in this two-time ensemble is that the state {\it will continue to resemble the low entropy boundary conditions.} The entire evolution is sensitive to the presence of {\it both} boundary conditions. In fact, as noted already, simple models of statistical systems with a boundary condition of low entropy at two times bear out these intuitively transparent expectations \cite{Cocke,Schulman2,Schulman3,Schulman4}. Moreover, in the case of Markov processes Schulman has demonstrated the described behaviour {\it analytically} \cite{Schulman5}. This confluence of intuitive clarity, and, for simple systems, analytic and (computerized) experimental evidence will be taken as suggestive that the significantly more complicated physical system considered in this work (matter fields in a dynamic universe) behaves in a qualitatively similar fashion when burdened with time symmetric boundary conditions.\footnote{Of course, the physics must at the completely fine-grained level be consistent with whatever dynamics the system obeys, including such constraints as conservation laws. As a moment's reflection on classical Hamiltonian systems with two-time phase space boundary conditions reveals, this may be a severe constraint! A purely stochastic dynamics (no dynamical conservation laws) allows systems great freedom to respect the RTH, and conclusions drawn from the behaviour of such systems may therefore be misleading. The restrictions on sensible boundary conditions in time neutral generalized quantum mechanics noted in the last paragraphs of section \ref{sec:motivations}\ (which arise essentially as a result of the requirements of decoherence) are examples of phenomena with no counterpart in stochastic systems. See also the discussion of gravitational collapse in section \ref{sec:collapse}.} The lesson of this section is that the place to look for signs of statistical time symmetry in our universe is in physical processes with long ``relaxation times," or more generally in any process which might couple the expanding and recollapsing eras \cite{TSA}. Such processes might be constrained by the presence of a final boundary condition. Obvious candidates include decays of long-lived metastable states \cite{Wheeler1,Wheeler2,Schulman3,Schulman5,TSA}, gravitational collapse \cite[see also \cite{Schulman3}]{Penrose1,Penrose2,Laflamme} and radiations \cite{DT,TSA} with great penetrating power such as neutrinos, gravitational waves, and possibly electromagnetic radiation. Thus, there are a variety of tests a cosmological model with both an initial and a final condition must pass in order to provide a plausible description of our universe. The next two sections discuss the extragalactic background radiation as an example of a physical prediction that is expected to be sensitive to the presence of a final boundary condition (time symmetrically related to the initial.) I focus on electromagnetic radiation because it is the most within our present observational and theoretical grasp, but the essence of the discussion is relevant to any similar wave phenomenon. Finally, in section \ref{sec:difficulties}\ I offer a few comments on some other issues that need to be addressed in any attempt to describe our universe by a model such as the one sketched in section \ref{sec:motivations}. \section{The Opacity of the Future Light Cone} \label{sec:opacity} \setcounter{equation}{0} The aim of this section is to extend arguments of Davies and Twamley \cite{DT} showing that a photon propagating in intergalactic space is likely to survive until the epoch of maximum expansion (assuming the universe to be closed), no matter how long the total lifetime of the universe. That is, the future light cone is essentially transparent over a wide range of frequencies for extinction processes relevant in the intergalactic medium (IGM), with an optical depth of at most $\tau \sim .01$ at optical frequencies. The physics in this result is that, in cases of physical interest, the dilution of scatterers due to the expansion of the universe wins out over the extremely long path length the photon must traverse. As a consequence, the integrated background of light from galaxies in the expanding phase will still be present in the recollapsing phase. (As explained in section \ref{sec:egbr}, it is this fact which implies that there is an ``excess" EGBR in a time symmetric universe that is not associated with galaxies to our past.) To show this, I compute, for a fairly general class of absorption coefficients, the optical depth between the present epoch and the moment of maximum expansion. For realistic intergalactic extinction processes this optical depth turns out to be small. Indeed, the opacity of the future light cone turns out to be dominated by ``collisions" of intergalactic photons with other galaxies, if they are regarded as completely opaque hard spheres. Even in the limit that the lifetime of the universe $T$ becomes infinite ($\Omega \rightarrow 1$ from above) all of the relevant processes yield finite optical depths. As this is essentially the limit of a flat universe, it is no surprise that extremely simple expressions result. The results of this section are modest extensions of the work of Davies and Twamley \cite{DT}. For the intergalactic extinction mechanisms and at the frequencies they consider, the formulae for the opacity derived here give numbers in agreement with their results (using the same data for the IGM, of course.) The present work is of slightly broader applicability in that the opacity is evaluated for a fairly general class of frequency dependent extinction coefficients (not just for a few specific processes), and its behaviour as the lifetime of the universe becomes arbitrarily long is determined. It turns out that quite generally, the asymptotic limit is in fact of the order of magnitude of the {\it upper} limit to the opacity in a closed universe. These results are less general than that of \cite{DT} in that I do not include the effects of a cosmological constant (which makes the perturbative analysis below significantly more awkward.) However, a small cosmological constant does not effect the qualitative nature of the conclusions of this section. \subsection{The Future Light Cone Can Be Transparent} \label{sec:transparent} Optical depth $\tau$ is defined by \begin{equation} d\tau = \Sigma \, dl, \end{equation} where $\Sigma$, the linear extinction coefficient, is the fractional loss of flux per unit (proper) length $l$, and is given microscopically by \begin{equation} \Sigma = \sigma n \label{eq:egbrsigmadef} \end{equation} for incoherent scattering from targets with cross section $\sigma$ and proper number density $n$ (this neglects stimulated emission and scattering into the line of sight.) Given $\tau$, the flux density along the line of sight thus obeys \begin{equation} i(l) = i_{0}e^{-\tau (l)}. \end{equation} Put another way, the probability a photon will propagate a distance $l$ without being absorbed is $ e^{-\tau (l)} $. For further details, see for example \cite{RL}. In order to compute $\tau$ we need $\Sigma (l)$. In the approximation (appropriate to the calculation of optical depths between the present and the moment of maximum expansion) that the universe is exactly described by closed, dust-filled Friedmann-Robertson-Walker, it turns out to be helpful to trade in the dependence on proper length for time. The metric is \begin{equation} ds^{2} = a^{2}\, [ -d\eta^{2} + d\Omega_{3}^{2} ] , \end{equation} where \begin{equation} d\Omega_{3}^{2} = d\chi^{2} + \sin^{2}\chi \, d\Omega_{2}^{2} \end{equation} is the metric on the unit 3-sphere, and \begin{equation} dt = a \, d\eta \end{equation} relates the cosmological time (proper time in the cosmological rest frame) to the conformal time $\eta $. For dust the time dependence of the scale factor $a$ can be expressed parametrically as \begin{equation} a(\eta) = M \, (1-\cos \eta ), \label{eq:a=} \end{equation} so that \begin{equation} t(\eta ) = M \, (\eta - \sin \eta ). \label{eq:t=} \end{equation} The lifetime of the universe is then $ T = 2\pi M$. Here $ M = \frac{4\pi}{3}\rho a^{3} $ is a constant as the universe expands, $\rho $ being the mass density of the dust. $M$ is related to more familiar cosmological parameters by \begin{equation} M = \frac{1}{H_{0}}\frac{q_{0}}{(2q_{0}-1)^{\frac{3}{2}}}. \label{eq:M} \end{equation} Employing the symmetry of the model to take a photon's path as radial, $ds^{2} = 0$ gives $dl = dt = a \, d\eta $, from which \begin{equation} d\tau = \Sigma a \, d\eta. \label{eq:dtau=} \end{equation} The high symmetry of Friedmann-Roberston-Walker also means that nearly all the relevant physical quantities simply scale as a power of $a$. Thus, as will be seen explicitly in the next subsection, it is necessary to consider only extinction coefficients of the form \begin{equation} \Sigma = \Sigma_{0} \, {\left( \frac{a_{0}}{a}\right)}^{p+1} , \label{eq:sigma=} \end{equation} where $a_{0}$ is some fiducial scale factor (conventionally the present one) and $p$ is a number. The goal is to compute the optical depth between now and the moment of maximum expansion. ``Now" will be taken to be the time $t_{0}$ from the big bang to the present. For absorption coefficients of the form (\ref{eq:sigma=}), the optical depth of the future light cone is \begin{eqnarray} \tau & = & \int_{\tau(t_{0})}^{\tau(T/2)}\! d\tau \nonumber \\ & = & \int_{\eta_{0}}^{\pi }\! \Sigma \, a \, d\eta \nonumber \\ & = & \Sigma_{0} \, t_0 \, g_{p}(\eta_{0}) \label{eq:tau=} \end{eqnarray} using (\ref{eq:a=}) and (\ref{eq:dtau=}). Here $\eta_{0}$ is the conformal time of the present epoch, and $g_{p}(\eta_{0})$ is the dimensionless function \begin{eqnarray} g_{p}(\eta_{0})& \equiv & \left( \frac{M}{t_{0}}\right) {\left( \frac{a_{0}}{M}\right)}^{p+1} \int_{\eta_{0}}^{\pi} \frac{d\eta }{(1-\cos \eta )^{p}} \nonumber \\ & = & \frac{(1-\cos \eta_{0} )^{p+1}}{\eta_{0}-\sin\eta_{0}} \int_{\eta_{0}}^{\pi} \frac{d\eta}{(1-\cos \eta )^{p}} . \label{eq:g=} \end{eqnarray} For integral and half-integral $p$ explicit evaluation of $g_{p}(\eta_{0})$ is possible, but not terribly illuminating. In the limit that the total lifetime of the universe $T$ is very long compared to $t_{0}$, however, simple expressions for any $p$ result. (This is no surprise as the results must approach those of a flat universe.) One straightforward procedure involves inverting (\ref{eq:t=}) to get a power series in ${\left(\frac{t_{0}}{M}\right)}^{\frac{1}{3}}$ for $\eta_{0}$, and using this to evaluate the asymptotic behaviour of $g_{p}(\eta_{0})$ as $M$ becomes large relative to $t_{0}$, which is held fixed. ($T=2\pi\, M \gg t_{0}$ corresponds to $\eta_{0} \ll 1$.) It is then tedious but straightforward to show that \begin{equation} g_{p} \sim \left\{ \begin{array}[c]{ll} \frac{3}{2p-1} \left[ 1+\frac{p+1}{10(2p-3)} {\left( \frac{6t_{0}}{M} \right) }^{\frac{2}{3}} \right] & \mbox{$ p > \frac{1}{2} \ ; \ p \not= \frac{3}{2} $} \\ \frac{3}{2} \left[ 1-\frac{1}{12}{\left( \frac{6t_{0}}{M} \right)}^{\frac{2}{3}} \ln\left(\frac{6t_{0}}{M}\right) \right] & \mbox{$p = \frac{3}{2}$}\\ \ln\left(\frac{M}{t_{0}}\right) \left[ 1-\frac{3}{40}{\left(\frac{6t_{0}}{M}\right) }^{\frac{2}{3}} \right] & \mbox{$p=\frac{1}{2} $}\\ 3\pi \frac{\Gamma (1-2p)}{\Gamma^{2} (1-p)} {\left( \frac{M}{6t_{0}}\right) }^{\frac{1-2p}{3}} - \frac{3}{1-2p} & \mbox{$-\frac{1}{2} < p < \frac{1}{2}$} \\ 12{\left( \frac{M}{6t_{0}}\right) }^{\frac{2}{3}} - \frac{9}{5} & \mbox{$ p = -\frac{1}{2}$} \\ 3\pi \frac{\Gamma (1-2p)}{\Gamma^{2} (1-p)} {\left(\frac{M}{6t_{0}}\right) }^{\frac{1-2p}{3}} \left[ 1-\frac{p+1}{20}{\left( \frac{6t_{0}}{M}\right) }^{\frac{2}{3}} \right] & \mbox{$ p < -\frac{1}{2}$} \end{array} \right. \label{eq:g} \end{equation} In each case only the leading order correction in $\frac{t_{0}}{M}$ has been retained. The most important thing to notice is that for $p > \frac{1}{2}$, $g_{p}$ is perfectly finite even as the lifetime of the universe becomes arbitrarily big, and as $\frac{t_{0}}{M}$ becomes very small the opacity converges to the value it would have in a flat universe. It is clear that for $p > \frac{3}{2}$ the $\Omega_{0} = 1$ result is a local lower limit on the opacity of the future light cone, as may be verified directly also from the available exact results. However, for reasonable $p$ the corrections to the flat universe result are only a factor of order one. (In fact, examination of the exact results reveals that the {\it maximum} of $g_{p}$ as one varies $\frac{t_{0}}{M}$ is at most $20\%$ larger than the flat universe result for $p \sim {\rm few}$. This is a good thing, because unless $\Omega$ is fairly close to one, $\frac{t_{0}}{M}$ is not a particularly small parameter! In terms of familiar cosmological parameters, \begin{equation} \frac{t_{0}}{M}= \left[ \cos^{-1}({q_{0}}^{-1}-1) -{q_{0}}^{-1}(2q_{0}-1)^{\frac{1}{2}}\right].) \label{eq:t/M} \end{equation} Physically, what's at work is the competition between the slower expansion rates of universes with larger $\Omega_{0}$'s, which tends to increase the opacity because the scattering medium isn't diluted as rapidly, and the decrease in the opacity due to the shortened time between the present epoch and the moment of maximum expansion.\footnote{It will be noticed by combining (\ref{eq:M}) and (\ref{eq:t/M}) that taking the limit $\Omega_{0} \rightarrow 1$ holding $t_{0}$ fixed requires $H_{0}$ to vary as well, converging to the flat universe relation $H_{0} = \frac{2}{3t_{0}}$. It is possible to repeat the entire analysis holding the observable quantity $H_{0}$ fixed instead of $t_{0}$ (for this purpose the more standard redshift representation is more useful than that in terms of conformal time used above), but unsurprisingly the conclusions are the same: the opacity is always finite for $p > \frac{1}{2}$; as $\Omega_{0}$ approaches one, the opacity approaches the flat universe result; and the {\it maximum} opacity for reasonable $p$ is only a factor $ \ltwid 1.2$ times the flat universe result. The resultant opacities are of course related in these limits via $t_{0} = \frac{2}{3H_{0}}$. Similarly, it is possible to perform a related analysis of more complicated extinction coefficients than (\ref{eq:sigma=}), for example incorporating the exponential behaviour encountered in free-free absorption (see (\ref{eq:Sff})) or in modeling evolving populations of scatterers with, for instance, a Schecter function type profile. However, these embellishments are not required in the sequel, and the techniques are tedious and fairly ordinary, so space will not be taken to describe them here.} To summarize, all of the processes relevant to extinction in the intergalactic medium have extinction coefficients that can be bounded above by a coefficient of the form (\ref{eq:sigma=}) with $p > \frac{1}{2}$. Using the limiting relationship $t_{0} = \frac{2}{3H_{0}}$, we have from (\ref{eq:tau=}) and (\ref{eq:g}) the simple result that for these processes, the upper limit to the opacity between the present epoch and the moment of maximum expansion, no matter how long the total lifetime of the universe, is of order \begin{equation} \tau = \frac{2}{2p-1}\frac{\Sigma_{0}\, c}{H_{0}}. \label{eq:tau} \end{equation} (I have returned to conventional units in this formula.) \subsection{The Opacity of the Future Light Cone} \label{sec:flc} In this section I apply the asymptotic formula (\ref{eq:tau}) for the upper limit to the optical depth of the FLC in a long-lived universe to show that if our universe is closed, photons escaping from the galaxy are (depending on their frequency) likely to survive into the recollapsing era. That is, the finite optical depths computed in the previous section are actually small for processes of interest in the intergalactic medium (IGM). For simplicity, I focus on photons softer than the ultraviolet at the present epoch; the cosmological redshift makes it necessary to consider absorption down to very low frequencies. It is important to note that in employing standard techniques for computing opacities the effects of the assumed statistical time symmetry of the universe are being neglected. As discussed in section \ref{sec:tsbc}\ and in section \ref{sec:difficulties}, when the universe is very large the thermodynamic and gravitational behaviour of matter will begin to deviate from that expected were the universe not time symmetric. Due to the manifold uncertainties involved here it is difficult to approach the effects of time symmetric boundary conditions on the opacity of the future light cone with clarity.\footnote{For example, how is scattering of light by a ``thermodynamically reversed" medium to be treated, as when light from the expanding era reaches the intergalactic medium in the recollapsing phase? The standard account assumes incoherent scattering. Thus a laser beam shone on a plasma is diffused. Time-reversing this description yields {\it extremely} coherent scattering from the plasma which reduces its entropy. Thus scattering or absorption of light correlated with sources (such as galaxies) in the expanding phase by material in the recollapsing phase appears to require entropy reducing (according to the observers of the recollapsing era) correlations in the matter there, in contradiction with the presumed local thermodynamic arrow (and with the RTH), in order to yield what there appears as emission. This is just the sort of detailed connection between the expanding and recollapsing eras which would lead one to expect physical predictions in a time symmetric model, even very near one of the boundary conditions, to be very different than those in a model with an initial condition only (section \ref{sec:tsbc}). This complication is closely related to the difficulty, mentioned in section \ref{sec:retardation}, in deriving the retardation of radiation in a universe which is time symmetric and in which the future light cone is transparent.} I shall assume they are not such as to increase it. This is reasonable as the dominant contribution to the opacity comes when the universe is smallest, where in spite of the noted complications the RTH is assumed to hold. What are the processes relevant to extinction of photons in the intergalactic medium? Because the IGM appears to consist in hot, diffuse electrons, and perhaps a little dust \cite{BFR}, extinction processes to include are Thomson scattering, inverse bremsstrahlung (free-free absorption), and absorption by dust. In addition, absorption by material in galaxies (treated as completely black in order to gauge an upper limit) is important. These processes will treated in turn. (A useful general reference on all these matters is \cite{RL}.) The conclusion will be that while absorption by galaxies and Thomson scattering are most significant above the radio, none of these processes pose a serious threat to a photon that escapes from our galaxy. This confirms the results of Davies and Twamley \cite{DT}, who however did not consider the possibly significant interactions with galaxies. Consequently I will be brief. Some results of Davies and Twamley regarding absorption mechanisms which may be important when the universe is very large and baryons have had time to decay are quoted at the end of this section. These do not appear to be significant either. (For high energy photons Compton scattering, pair production, photoelectric absorption by the apparently very small amounts of neutral intergalactic hydrogen, and interactions with CMBR photons will be important, but as none are significant below the ultraviolet I do not discuss them here. All can be treated by the same methods as the lower energy processes.) To begin, following Davies and Twamley \cite{DT}, I quote Barcons {\it et al.\ }\cite{BFR} on current beliefs regarding the state of the IGM in the form \begin{eqnarray} n_{{\rm H_{II}}} & = & \delta \, n_{0} \, {\left( \frac{a_{0}}{a} \right) }^{3} \nonumber \\ T_{{\rm H_{II}}} & = & \epsilon \, T_{0} \, {\left( \frac{a_{0}}{a} \right) }^{2} \label{eq:IGM} \end{eqnarray} where \begin{eqnarray} n_{0} & = & 1.12 \, h^{2} \, 10^{-7} \, {\rm\,cm}^{-3}, \ \ \delta \in (1,10) \nonumber \\ T_{0} & = & 10^{4} \, {\rm\,K} , \ \ \epsilon \in (1,10^{3}) \label{eq:IGMparams} \end{eqnarray} with the values $\delta = \epsilon = 1$ somewhat preferred by the authors. In addition, the present upper limit on a smoothly distributed component of neutral hydrogen is about $n_{{\rm H_{I}}} < 10^{-12} \, {\rm\,cm}^{-3}$. Thus, the intergalactic medium consists in hot (but non-relativistic) electrons, protons, and essentially no neutral hydrogen. The lack of distortions in the microwave background indicates its relative uniformity, at least to our past. From now on, $n$ and $T$ simply will be used to refer to the number density and temperature of intergalactic electrons. Finally, very little is known about a possible diffuse component of intergalactic dust \cite{BFR,WR}, except that there is probably very little of it. Most dust seems to be clumped around galaxies. Therefore I will ignore possible extinction due to it, subsuming it into the ``black galaxy" opacity. Davies and Twamley \cite{DT} make some estimates for one model for the dust, finding its contribution to the opacity insignificant. At any rate, models for the absorption coefficient due to dust \cite{Peebles,HW} all give a cross section $\sigma$ that falls with increasing wavelength, $\sigma \sim 1/{{\lambda}^{q}}$ with $1 \leq q \leq 4$, so that $\Sigma = \sigma \, n \propto {\left( \frac{a_{0}}{a}\right)}^{q+3}$ (neglecting of course a clumping factor expressing the fact that clumping decreases the opacity.) Thus $p_{{\rm dust}} = q + 2 > \frac{1}{2}$, the dust opacity is bound to be finite, and with a small present density of diffuse dust it is not surprising to find its contribution to be small. Before considering the optical depth due to interactions with intergalactic electrons, I will show that it is reasonable to approximate that most photons escaping our galaxy will travel freely through intergalactic space. That is, few photons will end up running into another galaxy. \subsubsection*{Collisions with Galaxies} \label{sec:collisions} Drastically overestimating the opacity due to galaxies by pretending that any photon which enters a galaxy or its halo will be absorbed by it (the ``black galaxy" approximation), and taking the number of galaxies to be constant, \begin{eqnarray} \label{eq:hardspheresigma} \Sigma_{{\rm gal}} & = & \sigma \, n \nonumber \\ & = & \sigma \, n_{0} \, {\left( \frac{a_{0}}{a} \right) }^{3}, \end{eqnarray} where $\sigma$ is the cross-sectional area of a typical galaxy and $n_{0}$ is their present number density. Thus from (\ref{eq:tau}), the upper limit on the opacity due to collisions with galaxies is \begin{equation} \tau = \frac{2}{3}\, \frac{\sigma \, n_{0} \, c}{H_{0}}. \end{equation} As noted above, this is finite (even as the lifetime of the universe becomes very large) because the dilution of targets due to the expansion of the universe is more important than the length of the path the photon must traverse. Notice that assuming target galaxies to be perfectly homogeneously distributed only overestimates their ``black galaxy" opacity. Volume increases faster than cross-sectional area, so clustering reduces the target area for a given density of material. As galaxy clustering is not insignificant today and will only increase up to the epoch of maximum expansion even in a time symmetric universe, the degree of overestimation is likely to be significant. Taking $n_{0}\sim .02\, h^{3}\, {\rm\,Mpc}^{-3}$, $\sigma = \pi\, r_{{\rm gal}}^{2}$ (where $r_{{\rm gal}} \sim 10^{4}\, h^{-1}\, {\rm\,pc}$), and $H_{0} \sim \frac{1}{3} \cdot 10^{-17}\, h\, {\rm\,s}^{-1}$ (here $.4 < h < 1 $ captures as usual the uncertainty in the Hubble constant) gives the upper limit \begin{equation} \tau \sim .01. \label{eq:HS} \end{equation} This can be interpreted as saying that at most about one percent of the lines of sight from our galaxy terminate on another galaxy before reaching the recollapsing era. By time symmetry, neither do most lines of sight connecting the present epoch to its time-reverse. \subsubsection*{Thomson Scattering} \label{sec:thomson} Use of the Thomson scattering cross section $ \sigma_{{\rm T}} = \frac{8\pi}{3}r_{0}^{2} = 6.65 \cdot 10^{-25}\, {\rm\,cm}^{-2}$ is acceptable for scattering from non-relativistic electrons for any photon softer than a hard X-ray ($\hbar \omega \ll mc^{2}$). Thus, for the frequencies I will consider, $\Sigma_{{\rm T}} = \sigma_{{\rm T}} n$ will suffice, giving \begin{eqnarray} \tau_{{\rm T}} & = & \frac{2}{3}\, \frac{\delta \sigma_{{\rm T}} n_{0} c}{H_{0}} \nonumber \\ & = & 4.7 (\delta h) 10^{-4}. \label{eq:T} \end{eqnarray} Recalling that $\delta$ is at worst one order of magnitude, it is clear that Thomson scattering is not signficant for intergalactic photons \cite{DT}. It is perhaps worth mentioning that quantum and relativistic effects only tend to decrease the cross section at higher energies. More significant for the purposes of this investigation is the observation that, at the considered range of frequencies, Thomson scattering does not change a photon's frequency, merely its direction. Thus Thomson scattering of a homogeneous and isotropic bath of radiation by a homogeneous and isotropic soup of electrons has {\it no effect} as regards the predictions of section \ref{sec:egbr}.\footnote{Were it significant, it would however be a means of hiding the {\it information} contained in the background.} \subsubsection*{Inverse Bremsstrahlung} \label{sec:bremsstrahlung} Even less significant than Thomson scattering for frequencies of interest is free-free absorption by the IGM \cite{DT}. From, {\it e.g.\ }\cite{RL}, the linear absorption coefficient for scattering from a thermal bath of ionized hydrogen is \begin{eqnarray} \Sigma_{{\rm ff}} & = & \frac{2e^{6}}{3m\hbar c} {\left( \frac{2}{3\pi km} \right) }^{\frac{1}{2}} n^{2}T^{-\frac{1}{2}} \nu^{-3} \overline{g}(b) (1-e^{-b}) \nonumber \\ &=&3.7\cdot 10^{8} n^{2} T^{-\frac{1}{2}} \nu^{-3} \overline{g}(b)(1-e^{-b}) {\rm\,cm}^{-1}. \label{eq:Sff} \end{eqnarray} in cgs units. Here $b \equiv \frac{h\nu}{kT}$, the factor $e^{-b}$ contains the effect of stimulated emission, and $\overline{g}(b)$ is a ``Gaunt factor" expressing quantum deviations from classical results. It is a monotonically decreasing function of $b$ which is of order one in the optical ({\it cf.\ }\cite{RL} for a general discussion and some references.) As \begin{eqnarray} b &=& \frac{h\nu}{kT} \nonumber \\ &=& \frac{h\nu_{0}}{kT_{0}} \left( \frac{a}{a_{0}}\right) \nonumber \end{eqnarray} increases as the universe expands, taking $\overline{g}(b) = g_{0}$, a constant of order one, will only overestimate the opacity. Similarly, following \cite{DT} in dropping the stimulated emission term will yield an upper limit to the free-free opacity. With $\epsilon = 1$, $\frac{h\nu_{0}}{kT_{0}} = 1$ when $\nu_{0} \sim 10^{14}\, {\rm\,s}^{-1}$, so stimulated emission will only lead to a noticeable reduction in $\Sigma_{{\rm ff}}$ well below the optical. (Actually, methods similar to that employed in section \ref{sec:transparent}\ can be employed to calculate this term, but as $\tau_{{\rm ff}}$ will turn out to be insignificant even neglecting it there is no need to go into that here.) With these approximations, $$ \Sigma_{{\rm ff}} \approx (4.6 \cdot 10^{-8}) g_{0}\delta^{2}\epsilon^{-\frac{1}{2}} h^{4} \nu_{0}^{-3} {\left( \frac{a_{0}}{a} \right) }^{2}, $$ and thus \begin{eqnarray} \tau_{{\rm ff}} & = & 2 \frac{\Sigma_{0} c}{H_{0}} \nonumber \\ &=&8.6\cdot 10^{20}h^{3} g_{0}\delta^{2}\epsilon^{-\frac{1}{2}} \nu_{0}^{-3}. \label{eq:ff} \end{eqnarray} Recalling that $\delta = \epsilon = 1 $ seem likely physical values, and noting that $\delta^{2}\epsilon^{-\frac{1}{2}} \ltwid 10^{2}$ at worst, taking $h^{3} g_{0}\delta^{2}\epsilon^{-\frac{1}{2}} = 1$ is not unreasonable for an order of magnitude estimate. Thus $\nu_{0} \sim 10^{7} {\rm s}^{-1}$ (long radio) is required to get $\tau_{{\rm ff}} \sim 1$. Since $\tau_{{\rm ff}} \propto \nu_{0}^{-3}$ it drops sharply for photons with present frequency above that. For instance, at 5000\AA $$ \tau_{{\rm ff}} = g_{0}\delta^{2}\epsilon^{-\frac{1}{2}} 10^{-24}, $$ and inverse bremsstrahlung is completely negligible. \subsubsection*{The Far Future} \label{sec:future} Finally, I mention that Davies and Twamley \cite{DT} consider what happens if baryons decay in a long lived universe. Following the considerations of \cite{PM}, they conclude that the positronium ``atoms" which will form far in the future (when the universe is large) remain transparent to photons with present frequencies in the optical. This is because the redshifted photons haven't enough energy to cause transitions between adjacent Ps energy levels. Similarly, if in the nearer future the electrons and protons in the IGM recombine to form more neutral hydrogen, this will also be transparent at the considered frequencies. \section[The EGBR in Time Symmetric Universe]{Extragalactic Background Radiation in a \protect\\ Statistically Time Symmetric Universe} \label{sec:egbr} \setcounter{equation}{0} \subsection{Lower Limit to the Excess Optical EGBR} \label{sec:limit} The goal of this section is to explain why, in a statistically time symmetric universe (such as one with the CPT-related boundary conditions discussed in section \ref{sec:motivations}), the optical extragalactic background radiation should be at least twice that expected in a universe which is not time symmetric, and possibly considerably more. Thus, assuming consistency with the RTH ({\it i.e.\ }the predictive assumption that physics near either boundary condition is practically insensitive to the presence of the other boundary condition, {\it cf.\ }section \ref{sec:tsbc}), it is possible to discover {\it experimentally} whether our universe is time symmetric. Section \ref{sec:observations}\ compares this prediction with present observations, concluding that the minimal prediction is consistent with upper limits on the observed optical EGBR. However, better observations and modeling may soon challenge even this minimal prediction. At optical wavelengths, the isotropic bath of radiation from sources outside our galaxy is believed to be due almost exclusively to galaxies on our past light cone \cite[are some good general references]{Peebles,Deep,IAU,EGBR}. There is no other physically plausible source for this radiation. In a model universe with time symmetric boundary conditions, however, there must in addition be a significant quantity of radiation correlated with the time-reversed galaxies which will exist in the recollapsing era, far to our future \cite{DT,TSA}. The reason for this is that light from our galaxies can propagate largely unabsorbed into the recollapsing phase no matter how close to open the universe is, as shown in \cite{DT} and in section \ref{sec:opacity}. This light will eventually arrive on galaxies in the recollapsing phase, or, depending on its frequency, be absorbed in the time-reversed equivalent of one of the many high column density clouds (Lyman-limit clouds and damped Lyman-$\alpha$ systems) present in our early universe \cite{Peebles,Deep,QSO}, in the intergalactic medium, or failing that, at the time-reversed equivalent of the surface of last scattering. This will appear to observers in the recontracting phase as emission by one of those sources sometime in their galaxy forming era. Since future galaxies, up to high time-reversed redshift, occupy only a small part of the sky seen by today's (on average) isotropically emitting galaxies, much of the light from the galaxies of the expanding phase will proceed past the recontracting era's galaxies. Thus most of this light will be absorbed in one of the other listed media. Because of the assumption of global homogeneity and isotropy, the light from the entire history of galaxies in the expanding phase will constitute an isotropic bath of radiation to observers at the time-reverse of the present epoch that is {\it in addition} to the light from the galaxies to {\it their} past. By time symmetry, there will be a similar contribution to our EGBR correlated with galaxies which will live in the recollapsing phase, over and above that due to galaxies on our past light cone. To us this radiation will appear to arise in isotropically distributed sources {\it other} than galaxies. This picture of a transparent, time symmetric universe is illustrated in figure 1. \begin{figure}[ht] \label{fig:egbr} \begin{center} \epsfig{file=egbrfig.eps \end{center} \caption{Schematic representation of the origin of the ``excess'' extragalactic background light correlated with the thermodynamically reversed galaxies of the recollapsing era. The model is a Friedmann-Robertson-Walker universe equipped with time symmetric boundary conditions requiring the universe to be smooth and in local thermodynamic equilibrium whenever it is small.} \end{figure} A lower limit to this excess background can be obtained by considering how much light galaxies to our past have emitted already ({\it cf.\ }section \ref{sec:observations}). According to observers at the time reverse of the present epoch, this background will (in the absence of interactions) retain its frequency spectrum and energy density because the size of the universe is the same. Thus, by time symmetry, at a {\it minimum} the predicted optical EGBR in a universe with time symmetric boundary conditions is twice that expected in a universe in which the thermodynamic arrow does not reverse. If much of the luminous matter in galaxies today will eventually be burned into radiation by processing in stars or galactic black holes, the total background radiation correlated with galaxies in the expanding phase could be several orders of magnitude larger, a precise prediction requiring a detailed understanding of the future course of galactic evolution \cite{DT}. Several other processes may also contribute significant excess backgrounds. These topics are discussed further below. \subsubsection*{The Prediction} A number of points in the summary argument above require amplification. First, however, I summarize the {\it minimal}\footnote{By ``minimal" I mean the lower limit in each band provided by taking the integrated background of light from galaxies in the expanding era to be only that which has {\it already} been emitted up to the present epoch. In the absence of absorption, by time symmetry this is the minimum background at the present epoch that must be correlated with galaxies in the recontracting phase, as in the previous paragraph.} predictions for the ``excess" extragalactic background ({\it i.e.\ }radiation from non-galactic sources to our past that is correlated with time-reversed galaxies) in bands for which the future light cone is transparent: \begin{itemize} \item isotropy: the ``future starlight" should appear in the comoving frame as an approximately isotropic background. This conclusion depends crucially on the assumed global validity of the cosmological principle. \item energy density: comparable to the present energy density in starlight due to the galaxies on our past light cone. This assumes the future light cone (FLC) is totally transparent. \item spectrum: similar to the present spectrum of the background starlight due to galaxies on our past light cone. Again, neglect of further emissions in the expanding phase makes this, by time symmetry, a lower limit in each band. This conclusion relies on the assumption of a transparent FLC in part to the extent that this implies a paucity of standard astrophysical mechanisms for distorting spectra. \end{itemize} Thus, at for instance optical frequencies, time symmetry requires an isotropic extragalactic background at least twice that due to galaxies on our past light cone alone.\footnote{To be totally accurate, the quantity of radiation absorbed or scattered into another band between the present epoch and its time reverse should be subtracted. However, the upper limit to the total FLC opacity (due to anticipated processes) computed in section \ref{sec:opacity}\ was of order $10^{-2}$, mostly due to a liberal (``black galaxy") assessment of the rate of interception of photons by galaxies, and I will therefore neglect such losses. Further, it is worth remembering that processes like Thomson scattering do not destroy photons or change their frequency, but only scatter them. Thus mere scattering processes may introduce isotropically distributed (via the cosmological principle) fluctuations in the background, but not change its total energy. Similarly, line or dust absorption usually result in re-radiation of the absorbed photons, conserving the total energy in an isotropic background (if the size of the universe doesn't change much before the photons are re-radiated), if not the number of photons with a given energy.} The potentially far greater background predicted (by time symmetry) if further emissions in the expanding phase are accounted for is a subject taken up in the sequel. \subsubsection*{Consistency with the RTH?} Before proceeding, a comment on the consistency of this picture is in order. As the ``excess" radiation is correlated with the detailed histories of future galaxies, the transparency of the future light cone does not appear consistent with the predictive assumption (the RTH) that physics in the expanding era should be essentially independent of the specifics of what happens in the recontracting phase. At a minimum, if the model is to be at all believable it is legitimate to demand that the required radiation appears to us to arise in sources in a fashion consistent with known, or at least plausible, astrophysics. Thus it may be that given a transparent FLC, the only viable picture of a time symmetric universe is one in which the radiation correlated with future galaxies {\it ``should" be there anyway, i.e.\ }be predicted also in some reasonable model of our universe which is {\it not} time symmetric, and consequently not be ``excess" radiation at all, but merely optical radiation arising in non-galactic sources during (or before) the galaxy forming era. On the basis of present knowledge this does not describe our universe. The presence of the radiation required by time symmetry and the transparency of the FLC appears to be in significant disagreement with what is known about our galaxy forming era, as will become apparent below. Were it the case for our universe that non-galactic sources provided a significant component of the optical EGBR, the difficulties with time symmetric boundary conditions would from a {\it practical} point of view be less severe. It is true that the non-galactic sources emitting the additional isotropic background would have to do so in just such a way that the radiation contain the correct spatial and spectral correlations to converge on future galaxies at the appropriate rate. This implies a distressingly detailed connection between the expanding and recontracting phases. However, if the emission rate and spectrum were close to that expected on the basis of conventional considerations these correlations (enforced by the time symmetric boundary conditions) would likely be wholly unobservable in practice, existing over regions that are not causally connected until radiation from them converges onto a future galaxy \cite{DT}, and thus not visible to local coarse grainings (observers) in the expanding era. In any event, the meaning of the transparency of the FLC is that starlight is by no means a ``short relaxation time process." Of course, on the basis of the models discussed in section \ref{sec:tsbc}, perhaps the conclusion to draw from this apparent inconsistency between the RTH and the transparency of the FLC should rather be, that physical histories would unfold in a fashion quite different from that in a universe in which $\rho_{\omega}=1$, namely, in such a way that such detailed correlations would never be required in the first place. The very formation of stars might be suppressed ({\it cf.\ }section \ref{sec:difficulties}). Be that as it may, to the extent that the universe to our past is well understood, there are {\it no} sources that could plausibly be responsible for an isotropic optical background comparable to that produced by galaxies. Such an additional background, required in a transparent, time symmetric universe, requires significant, observable deviations from established astrophysics. This is in direct contradiction with the RTH. Thus the entire structure in which an additional background is predicted appears to be both internally inconsistent (in that it is inconsistent with the postulate which allows predictions to be made in the first place), and, completely apart from observations of the EGBR discussed in section \ref{sec:observations}, inconsistent with what is known about our galaxy forming era. This should be taken as a strong argument that our universe does not possess time symmetric boundary conditions. Nevertheless, in order to arrive at this conclusion it is necessary to pursue the consequences of assuming the consistency of the model. The situation may be stated thus: {\it either} our universe is not time symmetric, {\it or} there is an unexpected contribution to the optical EGBR due to non-galactic sources to our past (and there are indeed detailed correlations between the expanding and recontracting eras), {\it or} perhaps our future light cone is not transparent after all. This latter possibility, perhaps related to the considerable uncertainty regarding the state of the universe when it is large, seems the last resort for a consistent time symmetric model of our universe. \subsection{Amplifications} \label{sec:amplifications} It is now appropriate to justify further some of the points made in arriving at the prediction of an excess contribution to the EGBR. Claims requiring elaboration include: i) most of the light from the expanding era's galaxies won't be absorbed by the galaxies of the recontracting phase, and {\it vice-versa}; ii) it will therefore be absorbed by something else, and this is inconsistent with the early universe as presently understood; iii) a detailed understanding of the old age and death of galaxies, as well as other processes when the universe is large, may lead to a predicted EGBR in a time symmetric universe that is orders of magnitude larger than the minimal prediction outlined above. I deal with these questions in turn. \subsubsection*{The ``Excess" EGBR is Not Associated with Galaxies to Our Past} A photon escaping our galaxy is unlikely to encounter another galaxy before it reaches the time reverse of the present epoch. In fact, as shown in section \ref{sec:opacity}, galaxies between the present epoch and its time reversed equivalent subtend, at most, roughly a mere $ 2 \times .01 = 2\% $ of the sky (\ref{eq:HS}) (neglecting curvature and clumping.) In light of the present lack of detailed information about our galaxy forming era (and via time symmetry its time-reverse), a photon's fate after that is more difficult to determine. A straightforward extrapolation of the results of section \ref{sec:opacity}\ (or {\it cf. e.g.\ }section 13 of Peebles \cite{Peebles}) shows that the optical depth for encounters with galaxies of the same size and numbers as today is $\tau \sim .01 (1+z)^{\frac{3}{2}}$ between a redshift of $z$ and today. Again this neglects curvature (hence overestimating $\tau$) and clumping (which now underestimates $\tau$.) Assuming the bright parts of galaxies form at $z \sim 5$, this gives only $\tau \sim .14$, and the sky isn't covered with them until $z \sim 20$. This however is roughly at the upper limit on how old galaxies are thought to be. On the other hand, examination of quasar spectra (out to $z \sim 5$) show that most lines of sight pass through many clouds of high column densities of hydrogen called Lyman-$\alpha$ forest clouds and, at higher densities, damped Lyman-$\alpha$ systems. (Peebles \cite{Peebles} is a useful entry point on all of these matters, as is \cite{Deep}. \cite{QSO} are the proceedings of a recent conference concerned with these Lyman systems.) The highest density clouds may be young galaxies, but if so galaxies were more diffuse in the past as the observed rate of interception of an arbitrary line of sight with these clouds is a factor of a few or more greater than that based on the assumption that galaxy sizes are constant. (Obviously, this would not be too surprising.) For instance, for the densest clouds Peebles \cite[section 23]{Peebles} relates the approximate formula $$ \frac{dN}{dz} = 0.3 \, {\Sigma_{20}}^{-0.46} $$ for the observed interception rate per unit redshift of a line of sight with a cloud of column density greater than or equal to $\Sigma_{20}$ (in units of $10^{20} {\rm\,cm}^{-2}$), in a range of redshifts about $z=3$. For Lyman-$\alpha$ forest clouds, $\Sigma \gtwid 10^{14} {\rm\,cm}^{-2}$, the interception rate is considerably higher. (For some models see \cite{Tytler,Sargent}.) Thus an arbitrary line of sight arriving on our galaxy from a redshift of five, say, is likely to have passed through at least one cloud of column density comparable to a galaxy, and certainly many clouds of lower density. What might this mean for time symmetry? (For specificity I shall concentrate on photons which are optical today, say around 5000\AA. This band was chosen because at these wavelengths we have the luxury of the coincidence of decent observations, relatively well understood theoretical predictions for the background due to galaxies, the absence of other plausible sources for significant contributions, and a respectable understanding of the intergalactic opacity, including in particular some confidence that the future light cone is transparent.) Photons at 5000\AA\ today are at the Lyman limit (912\AA) at $z \approx 4.5$, and so are ionizing before that. At these redshifts the bounds on the amount of smoothly distributed neutral hydrogen (determined by independent measures such as the Gunn-Peterson test \cite{Peebles}) are very low ({\it cf.\ }section \ref{sec:flc}), presumably because that part of the hydrogen formed at recombination which had not been swept into forming galaxies was ionized by their radiation. Before the galactic engines condensed and heated up, however, this neutral hydrogen would have been very opaque to ionizing radiation. Similarly, near $z \sim 4$ Lyman-limit clouds with $\Sigma \gtwid 10^{17} {\rm\,cm}^{-2}$ are opaque at these frequencies. The upshot is that most photons from our galaxies which are optical today will make it well past the time reverse of the present epoch, likely ending up in the (time-reversed) L-$\alpha$ forest or in a young (to time reversed observers!) galaxy by $\tilde{z} \sim 4$. (Here $\tilde{z}$ is the epoch corresponding to the time-reverse of redshift $z$.) The very few that survive longer must be absorbed in the sea of neutral hydrogen between $\tilde{z}=0$ and (their) recombination epoch, $\tilde{z} \sim 1000$. Now, the important point is that on average, galaxies radiate isotropically into the full $4\pi$ of sky available to them. The lesson of the previous paragraph is that most lines of sight from galaxies in the expanding phase will not encounter a high column density cloud until a fairly high time-reversed redshift, $\tilde{z} \sim {\rm few}$, at which point many lines of sight probably {\it will} intersect one of these proto-galaxies or their more diffuse halos. If most photons from our galaxies have not been absorbed by this point, {\it this is not consistent with time symmetry}: the rate of emission of (what is today) optical radiation by stars in galaxies could not be time symmetric if the light of the entire history of galaxies in the expanding phase ends up only on the galaxies of the recollapsing phase at high $\tilde{z}$ (due consideration of redshifting effects is implied, of course.) Put another way, time symmetry requires the specific energy density in the backround radiation to be time symmetric. Thus the emission rate in the expanding era must equal (what we would call) the absorption rate in the recontracting phase. If stars in galaxies were exclusively both the sources (in the expanding phase) and sinks (as we would call them in the recontracting phase) of this radiation, galactic luminosities in the expanding phase would have to track the falling rate of absorption due to photon ``collisions" with galaxies. This is absurd. At the present epoch, for example, {\it at all frequencies} galaxies would (by time symmetry) have to be absorbing the diffuse EGBR (a rate for which the upper limit is determined entirely by geometry in the ``black galaxy" approximation) at the same rate as their stars were radiating (a rate that, in a time symmetric universe which resembles our own, one expects to be mostly determined by conventional physics.)\footnote{This is illustrated in the appendix with a simplified model. Related considerations may be used to put detailed constraints on the self-consistency of time symmetry, but I do not address that any further beyond the appendix. The essential point has already been made.} That is, stars would be in radiative equilibrium with the sky! This may be called the ``no Olber's Paradox" argument against the notion that a single class of localized objects could be exclusively responsible for the EGBR in a transparent universe equipped with time symmetric boundary conditions. (It might be thought that this problem would be solved if the time symmetric boundary conditions lead galaxies to radiate preferentially in those directions in which future galaxies lie. This is not a viable solution, because as noted above, only a small fraction of the sky is subtended by future galaxies up to high time reversed redshift. The deviations from isotropic emission would be dramatic.) Thus, the option consistent with time symmetry is that most galactic photons which are optical today will ultimately be absorbed in the many (by time symmetry) time-reversed Lyman-$\alpha$ forest clouds or Lyman-limit clouds believed to dwell between galaxies, and not in the stars of the time-reversed galaxies themselves.\footnote{This may be disappointing. A nice picture of a time symmetric universe might have photons from our galaxies arriving at time-reversed galaxies in the recollapsing era, appearing as their emissions. Even ignoring the highly detailed correlations between the expanding and recollapsing phases this would imply, the scheme could only work if radiation could be removed by galaxies in the recollapsing phase at the same rate it is emitted in the expanding. As noted, for isotropically emitting sources this is forbidden by time symmetry of the emission rate and geometry.} Fortunately for the notion of time symmetry this indeed appears to be the case. Careful studies of the opacity associated with Lyman systems \cite{Zuo,Madau}, indicates, within the bounds of our rather limited knowledge, that the light cone between $z=4.5$ and $z=0$ is essentially totally opaque to radiation that is 5000\AA\ at $z=0$, and that this is due largely to Lyman clouds near $z \sim 4$ and in the middle range of observed column densities, $\Sigma \sim 10^{16-17} {\rm\,cm}^{-2}$ or so. (To be honest, it must be admitted that hard data on just such clouds is very limited \cite{MM,Madau}.) We have now arrived at a terrible conundrum for the notion of time symmetry. Even if one is willing to accept the amazingly detailed correlations between the expanding and recontracting eras that reconciling a transparent future light cone with time symmetry requires, and even if the ``excess" radiation correlated with the galaxies of the recollapsing era were to be observed, this picture is incompatible with what little is known about the physical properties of the Lyman-$\alpha$ forest. Recalling the minimal prediction above for the excess background required by time symmetry, the prediction is that the Lyman-$\alpha$ forest has produced an amount of radiation at least comparable to that produced by the galaxies to our past. There is no mechanism by which this is reasonable. {\it There is no energy source to provide this amount of radiation.} More prosaically, the hydrogen plasma in which the clouds largely consist is observed (via determination of the line shape, for example) to be at kinetic temperatures of order $10^{4-5}K$, heated by quasars and young galaxies \cite{Carswell,Giallongo}. Thermal bremsstrahlung is notoriously inefficient, and line radiation at these temperatures is certainly insufficient to compete with nuclear star burning in galaxies! At for instance 5000\AA\ today, essentially {\it no} radiation is expected from forest clouds at all, let alone an amount comparable to that generated by galaxies. Remembering that by redshifts of 4.5 the Lyman forest is essentially totally opaque shortward of 5000\AA\ (observed) \cite{Madau}, it might have been imagined that an early generation of galaxies veiled by the forest heated up the clouds sufficiently for them to re-radiate the isotropic background radiation required by time symmetry. While it is true that quasars and such are likely sources of heat for these clouds \cite[for example]{Giallongo}, aside from the considerable difficulties in getting the re-radiated spectrum to resemble that of galaxies, the observed temperatures of the clouds are entirely too low to be compatible with the {\it minimum} amount of energy emission in the bands required. (A related restriction arises from present day observations of cosmological metallicities, which constrain the amount of star burning allowed to our past. If observed discrete sources came close to accounting for the required quantity of heavy elements, the contribution of a class of objects veiled completely by the Lyman forest would be constrained irrespective of observations of the EGBR. However, at present direct galaxy counts only provide about 10\% of the current upper limits on the extragalactic background light \cite{Tyson} ({\it cf.\ }section \ref{sec:observations}), the rest conventionally thought to arise in unresolvable galactic sources. Consequently, correlating formation of the heavy elements with observed discrete sources does not at present provide a good test of time symmetry. At any rate, such a test is likely to be a less definitive constraint on time symmetry because it is possible that a portion of the radiation lighting the Lyman forest from high redshift is not due to star burning, but to accretion onto supermassive black holes at the centers of primordial galaxies. Thus the best observational test is the most direct one, comparison of the observed EGBR with the contribution expected from galaxies.) The possibility that somehow the excess radiation does {\it not} come from the Lyman-$\alpha$ forest, but somehow shines through from other isotropically distributed sources even further in the past, is hardly more appealing. Familiar physics tells us that the forest is totally opaque to radiation that is 5000\AA\ at $z=0$. The conclusion had better be that the universe is not time symmetric, rather than that time symmetry engineers a clear path only for those photons correlated with galaxies in the recollapsing epoch (and not, say, the light from quasars.) Moreover, even if that were the case, analagous difficulties apply to the vast sea of neutral hydrogen that existed after recombination, totally opaque to ionizing radiation, and again to the highly opaque plasma which constituted the universe {\it before} recombination. It is possible to conjure progressively more exotic scenarios which save time symmetry by placing the onus on very special boundary conditions which engineer such rescues, but this is not the way to do physics. The only {\it reasonable} way time symmetry could be rescued would be if it were discovered that for reasons unanticipated here, the future light cone were not transparent after all, thus obviating the need for an excess background radiation with all its attendant difficulties. Otherwise, it is more reasonable to conclude that a universe with time symmetric boundary conditions would not resemble the one in which we actually live. \subsubsection*{Beyond the Minimal Prediction} Now that we have seen what kind of trouble time symmetry can get into with only the {\it minimal} required excess background radiation, it is time to make the problems worse. The background radiation correlated with the galaxies of the recollapsing era was bounded from below, via time symmetry, by including only the radiation that has been emitted by the galaxies to our past. But as our stars continue to burn, if the future light cone is indeed transparent it is possible a great deal more radiation will survive into the recollapsing era \cite{DT,TSA}. How much more? To get an idea of what's possible it is necessary to know both what fraction of the baryons left in galaxies will be eventually be burned into starlight, and when. For a rough upper bound, assume that {\it all} of the matter in galaxies today, including the apparently substantial dark halos (determined by dynamical methods to contribute roughly $\Omega_{{\rm gal}} \sim .1 $), will eventually be burned into radiation. To get a rough lower bound, assume that only the observed luminous matter ($\Omega_{{\rm lum}} \sim .004$) will participate significantly, and that only a characteristic fraction of about 4\% of {\it that} will not end up in remnants (Jupiters, neutron stars, white dwarfs, brown dwarfs, black holes, {\it etc.}) To overestimate the energy density of this background at $\tilde{z} = 0$, assume that all of this energy is released in a sudden burst at some redshift $z_{e} (< 0)$. Then by time symmetry, further star burning will yield a background of radiation correlated with time-reversed galaxies (expressed as a fraction of the critical density and scaled to $z=0$) somewhere in the range $$ (1+z_{e})^{-1} 10^{-6} \ltwid \Omega_{{\rm burn}} \ltwid (1+z_{e})^{-1} 10^{-3} \label{Oburn}.$$ (Here I have used the fact that the mass fraction released in nuclear burning as electromagnetic radiation is .007.) When $(1+z_{e})^{-1} \sim 1$ the upper limit is two orders of magnitude more than is in the CMBR today and three orders of magnitude more than present observational upper limits on a diffuse optical extragalactic background ({\it cf.\ }section \ref{sec:observations}). The lower bound, however, is comparable to the amount of radiation that has already been emitted by galaxies. Thus if the lower bound obtains, the prediction for the optical EGBR in a time symmetric universe is only of order three times that due to the galaxies to our past (if the excess background inferred from continued star burning is not distributed over many decades in frequency, and if most of this burning occurs near $z(\tilde{z})=0$.) As will be seen in section \ref{sec:observations}, this may still be consistent with present observational upper limits. On the other hand, if something closer to the upper limit obtains this is a clear death blow to time symmetry. A more precise prediction is clearly of interest. This would entail acquiring a detailed understanding of further galactic evolution, integrating over future emissions with due attention to the epoch at which radiation of a given frequency is emitted. (Naturally, this is the same exercise one performs in estimating the EGBR due to galaxies to our past \cite{Deep}.) Some idea of the possible blueshift ($(1+z_{e})^{-1}$) involved comes from estimating how long it will take our galaxies to burn out. This should not be more than a factor of a few greater than the lifetime of the longest lived stars, so a reasonable ballpark figure is to assume that galaxies will live for only another ten billion years or so. For convenience, assume that galaxies will become dark by $ t = n t_{0}$ for some $n$, where $t_{0}$ is the present age of the universe. To overestimate the blueshift at this time, assume the universe is flat, so that $$ (1+z_{e})^{-1} = (t/t_{0})^{\frac{2}{3}} = n^{\frac{2}{3}} .$$ For reasonable $n$'s this does not amount to a large (in order of magnitude) transfer of energy to the radiation from cosmological recontraction. \subsubsection*{Additional Sources of ``Excess" EGBR and the Far Future} In a similar fashion to continued burning of our stars, any isotropic background produced to our future might by time symmetry be expected to imply an additional contribution to the EGBR in an appropriately blueshifted band. For instance, even if continued star burning does not (by time symmetry) yield a background in contradiction with observations of the EGBR, it is possible that accretion onto the supermassive black holes likely to form at the centers of many galaxies could ultimately yield a quantity of radiation dramatically in excess of that from star burning alone.\footnote{I owe this suggestion to R. Antonucci.} In the absence of detailed information about such possibilities it is perhaps sufficient to note that ignoring possible additional contributions leads to a lower limit on the EGBR correlated with sources in the recontracting era, and I will therefore not consider them. There is one worrying aspect, however. As discussed in some detail by Page and McKee \cite{PM} for an approximately $k=0$ universe, and commented on in a related context in section \ref{sec:collapse}, if baryons decay then considerable photons may be produced by for instance the pair annihilation of the resulting electrons and positrons. Should not this, by time symmetry, yield a further contribution to the EGBR? The answer may well be yes, but there is a possible mechanism which avoids this conclusion. Somehow, with CPT symmetric boundary conditions, the density of baryons must be CPT symmetric. Therefore either baryons do not decay, or they are re-created\footnote{Note that in the former case CPT-symmetry requires that the observed matter-antimatter asymmetry inside the horizon does not persist at larger scales. In the latter case, {\it if} matter dominates homogeneously in the expanding era, then antimatter must dominate homogeneously in the recollapsing phase. Further discussion of CP-violation in T- and CPT-symmetric universes may be found in \cite{TSA}.} in precisely correlated collisions. (In the absence of a final boundary condition, the interaction rate would be too low for (anti-)baryon recombination to occur naturally.) The latter (boundary condition enforced) possibility appears extraordinary, but if baryon decay occurs in a universe with CPT symmetric boundary conditions, it could be argued that the best electrons and photons for the job would be just those created during baryon decay in the expanding phase, thereby removing this photon background. The ``no Olber's Paradox" argument, that most of an isotropically emitting source's light must end up in some homogeneous medium, and not, if time symmetry is to be preserved, equivalent time-reversed point sources, may not apply here if matter is relatively homogeneously distributed when the universe is large. Baryon decay {\it might} smooth out inhomogeneities somewhat before the resulting electrons and positrons annihilate. (This requires the kinetic energies of the decay products to be comparable to the gravitational binding energy of the relevant inhomogeneity.) Then the picture is no longer necessarily of localized sources emitting into $4\pi$, but of a more homogeneous photon-producing background that might cover enough of the sky to more reasonably secrete the required correlations for reconstruction of more massive particles in the recontracting phase. Nevertheless the extreme awkwardness of this scenario is not encouraging. The former possibility, clearly more palatable, is that baryons do not decay significantly either because $\Omega$ is not so near one after all that they have time enough to do so, or because the presence of the final boundary condition suppresses it. Either way, in this (possibly desperate) picture there is no additional background due to decaying baryons. A very similar question relates to the enormous number of particles produced in the last stages of black hole evaporation. This time, however, the objection that our black holes cover only a small portion of the recontracting era's sky, and consequently their isotropic emissions could not do the job of forming the white holes of the recontracting era (black holes to observers there) time symmetrically, would seem to be forceful. Thus if the universe is indeed very long-lived, black hole evaporation may well require yet an additional observable background. This may not be such a serious difficulty if $\Omega $ is not very close to one, however, as the time scales for the evaporation of galactic-scale black holes are quite immense. Further discussion of black holes in time symmetric cosmologies may be had in \cite{Zeh3,Penrose1,Penrose2}. One last point regarding the predictions described in this section needs to be made. Clearly, a loose end which could dramatically change the conclusions is the condition of matter in the universe when it is very large. This is uncertain territory, not the least because that is the era in a statistically time symmetric universe when the thermodynamic arrow must begin rolling over. Neglecting this confusing complication (reasonable for some purposes as many interactions are most significant when the universe is small), there is not a great deal known about what the far future should look like \cite{Rees,Davies,Islam,BT,Dyson,PM}. The study of Page and McKee \cite{PM} gives the most detailed picture in the case of a flat universe. As mentioned at the end of section \ref{sec:flc}, Davies and Twamley \cite{DT} find from this work that interactions of optical (at $z=0$) backgrounds with the electrons produced by baryon decay do not appear to be significant, primarily due to their diffuseness. On the other hand, if supermassive black holes (or any large gravitational inhomogeneities) appear, interactions with them may induce anisotropies in the future starlight. However, clumping only decreases the probability a line of sight intersects a matter distribution. Therefore large overdensities probably never subtend enough solid angle to interfere with most lines of sight to the recollapsing era unless gravitational collapse proceeds to the point where it dramatically alters homogeneity and isotropy on the largest scales. Because collapse is rather strongly constrained by time symmetry ({\it cf.\ }section \ref{sec:collapse}) I will not consider this possibility. Thus, insofar as the prediction of an ``excess" background radiation correlated with galaxies in the recollapsing era is concerned, the state of the universe when it is large would does not obviously play a substantial role. Nevertheless, given the manifold difficulties cited, the sentiment expressed above is that the best hope a time symmetric model has of providing a realistic description of our universe is that some unforseen mechanism makes the future light cone opaque after all. \subsection*{Summary} \label{sec:opacitysummary} To summarize, because our future light cone is transparent, neglecting starburning to our future and considering only the contribution to the EGBR from stars in our past provides an estimate of the total EGBR correlated with galaxies in the recollapsing era that is actually a lower limit on it. As mechanisms for distorting the spectrum generically become less important as the universe expands (barring unforseen effects in the far future), it is reasonable to take models for the present EGBR due to stars in our past as a minimal estimate of the isotropic background of starlight that will make its way to the recollapsing era. By time symmetry we can expect that at the same scale factor in the recollapsing era similar (but time-reversed) conditions obtain. As argued above, by time symmetry and geometry this ``future starlight" must appear to us as an additional background emanating from homogeneously distributed sources to our past {\it other} than galaxies. Therefore, if the universe has time symmetric boundary conditions which (more or less) reproduce familiar physics when the universe is small, and our future light cone is transparent, the optical extragalactic background radiation should be at least twice that expected to be due to stars in our past alone, and possesses a similar spectrum. If a considerable portion of the matter presently in galaxies will be burned into radiation in our future, by time symmetry the expected background is potentially much larger, and observations of the EGBR may already be flatly incompatible with observations. Nevertheless, in the next section I shall be conservative and stick with the minimal prediction in order to see how it jibes with observations. \subsection{Models and Observations of the Optical EGBR} \label{sec:observations} At optical wavelengths, it is generally believed that the isotropic background of radiation from extragalactic sources is due entirely to the galaxies on our past light cone \cite{Deep,IAU,EGBR}. As shown in the previous section, if our universe is time symmetric there must in addition be a significant contribution correlated with the galaxies of the recollapsing era which arises, not in galaxies, but in some homogeneously distributed medium, say for instance the Lyman clouds. The apparent inconsistency of this prediction with what is known about the forest clouds has been discussed above, and may be taken as an argument that our universe is not time symmetric. In this section judgement will be suspended, and the prediction of an ``excess" EGBR at least comparable to, but over and above, that due to galaxies to our past will be compared with experiment. The conclusion will be that current data are still consistent with time symmetry {\it if} our galaxies will not, in the time left before they die, emit a quantity of radiation that is considerably greater than that which they already have. A useful resource on both the topics of this section is \cite{Deep}. Tyson \cite{Tyson} has computed how much of the optical extragalactic background is accounted for by resolvable galaxies, concluding that known discrete sources contribute \begin{equation} \label{eq:tyson} \nu i_{\nu} \sim 3 \cdot 10^{-6} {\rm\,erg}\, {\rm\,cm}^{-2}\, {\rm\,s}^{-1}\, {\rm\,ster}^{-1} \end{equation} at 4500\AA. However, because very distant galaxies contribute most of the background radiation it is believed that unresolvable sources provide a significant portion of the EGBR. At present it is not possible to directly identify this radiation as galactic in origin. However, as understanding of galactic evolution grows so does the ability to model the optical extragalactic background due to galaxies. These predictions naturally depend on the adopted evolutionary models, what classes of objects are considered, the cosmological model, and so on. As representative samples I quote the results of Code and Welch \cite{CW} for a flat universe in which all galaxies evolve, \begin{equation} \label{eq:CW} \nu i_{\nu}\sim 8\cdot 10^{-6} {\rm\,erg}\, {\rm\,cm}^{-2}\, {\rm\,s}^{-1}\, {\rm\,ster}^{-1} \end{equation} at 5000\AA, and of Cole {\it et al.\ }\cite{CTS} for a similar scenario, \begin{equation} \label{eq:Cole} \nu i_{\nu}\sim 3\cdot 10^{-6} {\rm\,erg}\, {\rm\,cm}^{-2}\, {\rm\,s}^{-1}\, {\rm\,ster}^{-1}, \end{equation} also at 5000\AA. These figures are to be compared with the results of (extraordinarily difficult) observations. As surveyed by Mattila \cite{Mattila}, they give at 5000\AA\ an upper limit of \begin{equation} \label{eq:mattila} \nu i_{\nu}\ltwid 2\cdot 10^{-5}{\rm\,erg}\, {\rm\,cm}^{-2}\, {\rm\,s}^{-1}\, {\rm\,ster}^{-1}. \end{equation} As far as I am aware, there has been no direct detection of an optical radiation background of extragalactic origin. This upper limit represents what is left after what can be accounted for in local sources is removed. Comparing these results, it is clear that if current models of galactic evolution are reliable, present observations of the extragalactic background radiation leave room for a contribution from non-galactic sources that is comparable to the galactic contribution, but not a great deal more. These observations therefore constrain the possibility that our universe is time symmetric. If believable models indicate that further galactic emissions compete with what has been emitted so far, time symmetry could already be incompatible with experiment. A direct detection of the extragalactic background radiation, or even just a better upper limit, could rule out time symmetry on {\it experimental} grounds soon.\footnote{To be more careful, if direct HST galaxy counts, or reliable models of the (extra-)galactic contribution closely agree with the observations, the justifiable conclusion is that {\it if} our universe is time symmetric, then for some unknown reason our future light cone is not transparent.} (The ideal observational situation would be convergence of all-sky photometry and direct HST galaxy counts, allowing one to dispense with models completely.) \section{Further Difficulties with Time Symmetry} \label{sec:difficulties} \setcounter{equation}{0} In this section I comment on some issues of a more theoretical nature which must be faced in any attempt to construct a believable model of a time symmetric universe. Among other questions, in such a universe the difficulties with incorporating realistic gravitational collapse, and in deriving the fact that radiation is retarded, appear to be considerable. \subsection{Gravitational Collapse} \label{sec:collapse} A careful account of the growth of gravitational inhomogeneities from the very smooth conditions when the universe was small is clearly of fundamental importance in any model of the universe, time symmetric or not, not the least because it appears to be the essential origin of the thermodynamic arrow of time\footnote{For examples of concrete calculations connecting the growth of gravitational inhomogeneities with the emergence of a thermodynamic arrow see \cite{HaH,Hawking1,HLL}.} from which the other apparent time arrows are thought to flow \cite{PTA,Zeh2}. However, if matter in a closed universe is to be smooth at both epochs of small scale factor then it is incumbent to demonstrate that Einstein's equations admit solutions in which an initially smooth universe can develop interesting (non-linear) inhomogeneities such as galaxies which eventually smooth out again as the universe recollapses. This is because the universe is certainly in a quasiclassical domain now, and if it is assumed to remain so whenever the scale factor is large classical general relativity must apply. Laflamme \cite{Laflamme} has shown that in the linear regime there are essentially no perturbations which are regular, small, and growing away from both ends of a closed FRW universe, so that in order to be small at both ends a linear perturbation must be too small to ever become non-linear. But that is not really the point. Interesting perturbations must go non-linear, and there is still no proof of which I am aware that perturbations which go non-linear cannot be matched in the non-linear regime so as to allow solutions which are small near both singularities. Put differently, what is required is something like a proof that Weyl curvature must increase ({\it cf.\ }\cite{Penrose1,Penrose2}), {\it i.e.\ }that given some suitable energy conditions the evolution of gravitational inhomogeneity must be monotonic even in the absence of trapped surfaces, except possibly on a set of initial data of measure zero. While perhaps plausible given the attractive nature of gravity, proof has not been forthcoming. (It is important for present purposes that genericity conditions on the initial data are not a central part of such a proof, for physically realistic solutions which meet the time symmetric boundary conditions must describe processes in the recollapsing era which from our point of view would look like galaxies disassembling themselves. Reducing such a solution to data on one spacelike hypersurface at, say, maximum expansion, said data will be highly specialized relative to solutions with galaxies which do not behave so unfamiliarly. If such solutions with physically interesting inhomogeneities do exist, the real question here is whether they are generic {\it according to the measure defined by the two-time boundary conditions.} Since it ought to be possible to treat this problem classically, in principle this measure is straightforward to construct. The practical difficulty arises in evaluating the generic behaviour of solutions once they go non-linear. My own view is that it is highly likely that such solutions remain exceedingly improbable according to a measure defined by a {\it generic} set of (statistical) boundary conditions requiring the universe to be smooth when small. As noted, Laflamme \cite{Laflamme} has already shown that when the initial and final states are required to be smooth, the growth of inhomogeneity is suppressed if only linear perturbations are considered. Unless boundary conditions with very special correlations built in are imposed, probable solutions should resemble their smooth initial and final states throughout the course of their evolution, never developing physically interesting inhomogeneities.\footnote{One possible out is Schulman's observation that systems which exhibit chaos, as general relativity does, may be less restricted in the varieties of their behaviour by boundary conditions at two times than are systems with linear dynamics \cite{Schulman3}. This substantially unstudied possibility would obviously never emerge from Laflamme's linear analysis.} Note the concern here is not with occurrences which would be deemed unlikely in a universe with an initial condition only, but occur in a time symmetric universe because of the ``fate" represented by the final boundary condition, but with occurrences which are unlikely {\it even in a universe with (generic) time symmetric boundary conditions.}) A related question concerns collapsed objects in a time symmetric universe. Page and McKee \cite{PM} have studied the far future of a $k=0$ FRW universe under the assumption that baryons decay but electrons are stable. Assuming insensitivity to a final boundary condition their conclusions should have some relevance to the period before maximum expansion in the only slightly over-closed (and hence very long-lived) model universes that have mostly been considered here. As discussed in section \ref{sec:amplifications}, if the universe is very long lived it might be imagined that the decay of baryons and subsequent annihilation of the produced electrons and positrons could smooth out inhomogeneities, and also tend to destroy detailed information about the gravitational history of the expanding phase (by eliminating compact objects such as neutron stars, for instance.) Thus, even though interactions are unlikely to thermalize matter and radiation when the universe is very large \cite{PM} ({\it cf.\ }section \ref{sec:opacity}), there may be an analagous information loss via the quantum decay of baryons which could serve a similar function. (For a completely different idea about why quantum mechanics may effectively decouple the expanding and recollapsing eras, see \cite{Zeh1,Zeh2,Zeh3,KZ}.) In any case, if there is no mechanism to eliminate collapsed objects before the time of maximum expansion, then the collapsed objects of the expanding phase are the same as the collapsed objects of the recontracting phase, implying detailed correlations between the expanding and recontracting era histories of these objects which might lead to difficulties of consistency with the RTH. A particular complication is that it is fairly certain that black holes exist, and that more will form as inhomogeneity grows. The only way to eliminate a black hole is to allow it to evaporate, yet the estimates of Page and Mckee indicate that it is more likely for black holes to coalesce into ever bigger holes unless for some reason (a final boundary condition?) there is a maximum mass to the black holes which form, in which case they may have time enough to evaporate (though this requires $\Omega$ to be {\it exceedingly} close to one.) In fact, it may be imperative for a time symmetric scenario that black holes evaporate, else somehow they would have to turn into the white holes of the recollapsing era (black holes to observers there.) This is because we do not observe white holes today \cite{Penrose1,Penrose2}. (A related observation is that in order for the universe to be smooth whenever it is small, black/white hole singularities cannot arise \cite{Zeh3}.) Here the evaporation of black holes before maximum expansion would be enforced by the time symmetric boundary conditions selecting out only those histories for which there are no white holes in the expanding era and {\it mutatis mutandis} for the recollapsing era.\footnote{For more on black holes in time symmetric universes, see the discussion of Zeh \cite{Zeh3}.} On the other hand, if evaporating black holes leave remnants they too must be worked into the picture. Again, if the results of the stochastic models with two-time low entropy boundary conditions discussed above are to be taken seriously, the conclusion should probably be that boundary conditions requiring homogeneity when the universe is small suppress histories in which significant gravitational collapse occurs by assigning low probabilities to histories with fluctuations that will go non-linear. It hardly needs emphasizing that all of these considerations are tentative, and greater clarity would be welcome. \subsection{The Retardation of Radiation} \label{sec:retardation} Besides gravitational considerations, radiation which connects the expanding and recollapsing eras provides another example of a physical process which samples conditions near both boundary conditions \cite{DT}. While gravitational radiation and neutrinos are highly penetrating and are likely to provide such a bridge, in neither case are we yet capable of both effectively observing, and accurately predicting, what is expected from sources to our past. Therefore the primary focus of this investigation has been electromagnetic radiation. Above it was confirmed that modulo the obviously substantial uncertainty regarding the condition of the universe when it is large, even electromagnetic radiation is likely to penetrate to the recollapsing era. Section \ref{sec:egbr}\ was concerned with one relatively prosaic consequence of this prediction if our universe possesses time symmetric boundary conditions. Here I comment briefly on another. Maxwell's equations, the dynamical laws governing electromagnetic radiation, are time symmetric. It is generally believed that the manifest asymmetry in time of radiation {\it phenomena}, that is, that (in the absence of source-free fields) observations are described by retarded solutions rather than advanced, is ascribable fundamentally to the thermodynamic arrow of time without additional hypotheses. (For a contemporary review see \cite{Zeh2}.) However, if our universe possesses time symmetric boundary conditions then near the big bang the thermodynamic arrow of entropy increase runs oppositely to that near the big crunch. Since radiation can connect the expanding and recollapsing eras, the past light cone of an accelerating charge in the expanding era ends up in matter for which the entropy is increasing, while its future light cone terminates in matter for which entropy is supposed to be {\it decreasing}. If the charge radiates into its future light cone this implies detailed correlations in this matter with the motion of the charge which are incompatible with the supposed entropy decrease there (entropy increase to time reversed observers), although it is true that these correlations are causally disconnected to time-reversed observers, and consequently invisible to {\it local} coarse grainings defining a local notion of entropy for such observers. This state of affairs makes it difficult to decide whether radiation from an accelerating charge (if its radiation can escape into intergalactic space) should be retarded (from the perspective of observers in the expanding era) or advanced or some mixture of the two. The conditions under which the radiation arrow is usually derived from the thermodynamic arrow of surrounding matter do not hold. (Notice how this situation is reminiscent of the requirements necessary to derive retardation of radiation in the Wheeler-Feynman ``absorber theory" of electrodynamic phenomena \cite{WF}.) Hence the ability of radiation to connect the expanding and recollapsing epochs brings into question the self-consistency of assuming time symmetric boundary conditions on our universe together with ``physics as usual" (here meaning radiation which would be described as retarded by observers in both the expanding and recollapsing eras) near either end. The retardation of radiation is another important example of a physical prediction which would be expected to be very different in a universe with time symmetric boundary conditions than in one without. Once again, if the results of the simple stochastic models are generally applicable, the retardation of radiation should no longer be a prediction in such a universe.\footnote{Analytical elucidation of this idea is hoped to be the subject of a (far) future paper.} To summarize this section, consideration of gravitational collapse and radiation phenomena reveals that construction of a model universe with time symmetric boundary conditions which resembles our own may be a difficult task indeed. There are strong suggestions that a model with time symmetric boundary conditions which mimic our own early universe would behave nothing like the universe in which we live. Such a model would most likely predict a universe which remained smooth throughout the course of its evolution, with coupled matter components consequently remaining in the quasi-static ``equilibrium" appropriate to a dynamic universe. \section{Summation} \label{sec:summation} \setcounter{equation}{0} In spite of the oft-expressed intuitive misgivings regarding the possibility that our universe might be time symmetric \cite[for example] {Weinberg,PTA,Penrose1}, it has generally been felt that if sufficiently long-lived, there might be no way to tell the difference between a time symmetric and time asymmetric universe. Building on suggestions of Cocke, Schulman, Davies, and Gell-Mann and Hartle (among others), this work has explored in some detail one physical process which, happily, belies this feeling: no matter how long our universe will live, the time symmetry of the universe implies that the extragalactic background radiation be {\it at least} twice that due to the galaxies to our past. This is essentially due to the fact that light can propagate unabsorbed from the present epoch all the way to the recollapsing era. Moreover, geometry and time symmetry requires this ``excess" EGBR to be associated with sources {\it other} than the stars in those galaxies, sources which, according to present knowledge about the era during which galaxies formed, are not capable of producing this radiation! Thus the time symmetry of a closed universe is a property which is {\it directly accessible to experiment} (present observations are nearly capable of performing this test), as well as extremely difficult to model convincingly on the basis of known astrophysics. In addition, the other theoretical obstacles remarked upon briefly in sections \ref{sec:amplifications}\ and \ref{sec:difficulties}\ make it difficult to see how a plausible time symmetric model for the observed universe might be constructed. In particular, any such attempt must demonstrate that in a universe that is smooth whenever it is small, gravitational collapse can procede to an interesting degree of inhomogeneity when the universe is larger \cite{Laflamme}. This is necessary in order that the universe display a thermodynamic arrow (consistently defined across spacelike slices) which naturally reverses itself as the universe begins to recollapse. Furthermore, it appears unlikely that the usual derivation of the retardation of radiation will follow through in a time symmetric universe in which radiation can connect regions displaying opposed thermodynamic arrows. Finally, in the context of the time neutral generalized quantum mechanics employed as the framework for this discussion, unless the locally observed matter-antimatter asymmetry extends globally across the present universe,\footnote{Recall that this requires that the universe live long enough for nearly all baryons to decay, and reform into the antibaryons of the recollapsing era. This presents serious additional difficulties, {\it cf.\ }section \ref{sec:amplifications}.} natural choices of CPT-related boundary conditions yield a theory with trivial dynamics if the deviations from exact homogeneity and isotropy are specified in a CPT invariant fashion (see the last paragraph of section \ref{sec:motivations}). In sum, were the ``excess" EGBR which has been the primary concern of this investigation to be observed, it would appear necessary to place the onus of explanation of the fact that the final boundary condition is otherwise practically invisible upon very specially chosen boundary conditions which encode the details of physics in our universe. This would make it difficult to understand these boundary conditions in a natural way. On the dual grounds of theory and experiment, it therefore appears unlikely that we live in a time symmetric universe. (A definitive expurgation must await more thorough investigation of at least some of the aforementioned difficulties.)
\section{--~ Introduction} The CHARM2000 workshop\cite{CHARM2000} suggested that the goal for a future experiment be a factor $\sim10^2$ increase in statistics over the coming round of fixed-target charm experiments at Fermilab (E781 and E831). We consider the physics goals of such an ultrahigh-statistics charm experiment ($ \sim $ 100 million reconstuctable $ D $'s ). Some measurements will test the Standard Model, some will measure its parameters, and some will elucidate heavy-quark phenomenology. We can outline the major goals as follows: \begin{enumerate} \item Measurements which search for new physics\cite{Pakvasa} \begin{itemize} \parskip=0pt \item $ D^0 - \overline D {}^0$ mixing \item explicit flavor-changing neutral currents \item direct {\it CP} violation. \end{itemize} \item Measurements which test the heavy-quark symmetry of QCD in the charm sector \begin{itemize} \parskip=0pt \item form factors of semileptonic decays of charmed mesons \item masses and widths of orbitally-excited charmed mesons. \end{itemize} These measurements are key to extracting fundamental parameters from future beauty experiments. \item Measurements which probe aspects of perturbative and nonperturbative QCD (including higher-twist effects) \begin{itemize} \parskip=0pt \item nonleptonic singly and doubly Cabibbo-suppressed decay rates \item dynamics of charm hadroproduction. \end{itemize} \end{enumerate} \noindent We next discuss these topics in more detail. \section{--~ Searches for New Physics} \subsection{\underline {$ D^0 - \overline D {}^0 $ mixing}} $ D^0 - \overline D {}^0 $ mixing is one of the most interesting places to look for physics beyond the Standard Model. While Wolfenstein\cite{Wolfenstein} once suggested large long-distance or dispersive contributions to $ \Delta m_D $ within the Standard Model, more detailed calculations by Donoghue {\it et al.}\cite{Donoghue} give $ | \Delta m_D | \approx 10^{-6} $\,eV. Recent analyses\cite{Georgi,Ohl} based on heavy-quark effective theory (HQET) suggest that cancellations lead to $ | \Delta m_D | < 3.5 \times 10^{-8} $\,eV. Values of $ | \Delta m_D | $ as large as $ 10^{-4} $\,eV are possible in many models beyond the Standard Model\cite{EHLQ}~$^-$\cite{HW}. Thus there is a large window for observing new physics via $ D^0 - \overline D{}^0 $ mixing. A particularly intriguing example is discussed in a recent paper by Hall and Weinberg\cite{HW}, which emphasizes that electroweak theories with several scalar doublets are consistent with all known physics (especially with the approximate {\it CP} symmetry in the neutral-kaon sector), provide an alternative mechanism for {\it CP} violation, and have various interesting phenomenological features. In such models, tree-level scalar-exchange contributions to neutral $ K $- and $ B $-meson mass mixing are at about the level observed by experiments if Higgs bosons have masses in the 700\,GeV range. Hall and Weinberg say that ``although this means that little can be learned about the CKM matrix from $ \Delta m_K $ and $ \Delta m_B $, the case of $ D - \overline D $ [mixing] presents different opportunities. $ \ldots $ If we take the typical Higgs boson mass as near 1\,TeV to account for the observed values of $ | \Delta m_K | $ and $ | \Delta m_B | $, then the predicted value of $ | \Delta m_D | $ is close to the current experimental limit, $ | \Delta m_D | < 1.3 \times 10^{-4} $\,eV." CLEO II has reported a $ D^0 \rightarrow K^+ \pi^- $ signal with a branching ratio about $ 2 \times \tan^4 \theta_C \times B ( D^0 \rightarrow K^- \pi^+) $\cite{Cinabro}. This ``wrong-sign" kaon is a signature either of mixing or of doubly Cabibbo-suppressed decay (DCSD). The two are most easily separated in a fixed-target experiment, in which the lifetime of the $ D $ can be directly measured: a DCSD signal decays exponentially, while a mixing signal has an additional $ t^2 $ dependence, so that it peaks at $2\tau_D $. If the CLEO II signal is primarily a DCSD signal, then it presents an inescapable background for mixing studies using hadronic final states. Assuming this is the case, an experiment such as we consider should be sensitive to a mixing signal on top of a DCSD signal with $ | \Delta m_D | \approx2 \times 10^{-5} $\,eV. Morrison\cite{Morrison} has pointed out that similar sensitivity might be achievable also in semileptonic decays, which are free of the confounding effects of DCSD. Liu's thorough treatment\cite{Liu} includes the intriguing suggestion that for mixing arising from the decay-rate difference between the {\em CP} eigenstates $D_1$ and $D_2$, sensitivity an order of magnitude better might be achievable in singly Cabibbo-suppressed modes, by using the interference between mixing and DCSD to enhance the mixing signal. \subsection{\underline {Charm-changing neutral currents}} Some of the models cited above\cite{Buchmuller,Joshipura} also allow the possibility of explicit flavor-changing neutral currents (FCNC) in $ D $ decays. E791 has reported\cite{Nguyen} the best 90\%-confidence-level upper limit for the branching ratio of $ D^+ \rightarrow \pi^+ \mu^+ \mu^- $ of $ 1.3 \times 10^{-5}$. An ultrahigh-statistics experiment with good lepton identification would have a sensitivity to this and other FCNC decays (and also to lepton-number-violating decays) one to two orders of magnitude lower. \subsection{\underline {$C\!P$ violation}} The Standard Model predicts that direct {\it CP} violation (observed as the fractional difference between decay rates of particle and antiparticle to charge-conjugate final states) will be of order $10^{-3}$ or less in singly Cabibbo-suppressed $ D $ decays\cite{Bucella}. (In the Standard Model, {\it CP} should be an exact symmetry for Cabibbo-allowed and DCSD decays.) Physics beyond the Standard Model might contribute {\it CP}-violating amplitudes to decay rates, and there is a large window for observing new physics. At the level of statistics we consider here, sensitivity to {\it CP} asymmetries at the fraction-of-a-percent level in singly Cabibbo-suppressed decays and at the few-percent level in doubly Cabibbo-suppressed decays may be possible. Holding systematic uncertainties to the percent level will be challenging, and experimenters planning to make such measurements must consider carefully how systematic errors will be minimized and how they will be measured. \section{--~ HQET and Semileptonic Decays} \subsection{\underline {Testing HQET via orbitally-excited charmed mesons}} Within the Standard Model, it is generally agreed that heavy-quark symmetry can be used to predict many nonperturbative properties of hadrons containing a single heavy quark, and the most important of these predictions are for exclusive semileptonic $ B $-meson decays\cite{Wisgur1}. These nonperturbative effects will be important in extracting $ V_{ub} $ and $ V_{cb} $ from measured decay rates. Heavy-quark symmetry also relates the masses and widths of the orbitally-excited $ D^{**} $ mesons (including the $ D_s^{**} $ mesons), as has been discussed recently in papers by Isgur and Wise\cite{Wisgur2}, Ming-Lu, Isgur, and Wise\cite{Wisgur3}, and Eichten, Hill, and Quigg\cite{EHQ}. While some authors argue that the charm-quark mass is sufficiently large for the limit $ m_c \rightarrow \infty $ to be a good approximation\cite{EHQ}, others\cite{Godfrey} have argued that even for $ B $ mesons the $ m_Q \rightarrow \infty $ limit has not been reached. The experiment we consider will measure the masses and widths of the orbitally-excited $ D^{**} $ mesons with sufficient precision to confront theoretical models quantitatively. Where E691\cite{E691}, ARGUS\cite{ARGUS}, CLEO II\cite{CLEO}, and E687\cite{E687} have measured the $ D_2^{*0} $, $ D_2^{*+} $, and $ D_1^0 $ widths with 50\% fractional errors, such an experiment should be able to achieve few-percent fractional errors. To untangle the states and reflections which lie on top of each other, it will also be necessary to measure $ \pi^0 $'s and (perhaps) single photons well. However, the benefit of making these measurements is that they will establish how well heavy-quark symmetry works for charm and give theorists the numbers they need to develop a more complete phenomenology of $ B $ physics. \subsection{\underline {Semileptonic form factors}} High-statistics charm experiments will also contribute to our understanding of the form factors and helicity amplitudes of the vector mesons which can appear as decay products in both $ D $ and $ B $ decays. Extracting {\it CP}-violation parameters from measurements of branching ratios for decays such as $ B_d \rightarrow \rho^0 \psi $ and $ B_d \rightarrow K^* \psi $, which Dunietz\cite{Dunietz} advocates as the best place to measure the unitarity-triangle angle $ \gamma $, requires the best possible measurement of the $ \rho^0$ and $ K^* $ helicity amplitudes and form factors in the $ D $ semileptonic decays $ D^+ \rightarrow \rho^0 l \nu $ and $ D^+ \rightarrow K^{*0} l \nu $, as they should be the same in $ D $ as in $B $ decay. Assuming single-pole forms for the form factors, the mass of the pole should be measurable with better than 1\% precision. In $ D^+ \rightarrow K^{*0} l \nu $, it should be possible to measure the polarization of the $ K^* $, \begin{equation} {\Gamma_L / \Gamma_T} = {{ \int P^*_V t | H_0(t) | ^2 dt} \over {\int P^*_Vt \left [ | H_+(t) |^2 + |H_- (t) |^2 \right ] dt} } \end{equation} (the ratio of longitudinal to transverse form-factors), with percent statistical and systematic uncertainties. It should be possible to measure the polarization of the $ \rho^0 $ in the Cabibbo-suppressed decay with few-percent statistical accuracy. $ D_S \rightarrow \phi l \nu $ should be measured with similar precision, which will provide another test of the applicability of heavy-quark symmetry to the study of semileptonic decays. \subsection{\underline {Studying the CKM matrix with semileptonic decays}} Studying semileptonic decays also contributes directly to our knowledge of the CKM matrix. High-statistics charm experiments are able to measure the magnitudes of $ V_{cs} $ and $ V_{cd} $ from the semileptonic decays of the $ D $ mesons. The absolute decay rates depend on various well-measured constants (such as the $ D $ masses and lifetimes), the CKM matrix elements, and the form factors of the hadrons produced along with the leptons. Currently, $| V_{cd}| $ and $ | V_{cs} |$ are known with $ \pm 8 \% $ and $ \pm 20 \% $ precision\cite{PDG,Rosner}. {}From the branching ratios for the semileptonic decays $ D^0 \rightarrow \pi^- l^+ \nu_{l} $ and $ D^0 \rightarrow K^- l^+ \nu_{l} $, the ratio $ | V_{cd} | / | V_{cs} | $ should be determined with a statistical accuracy of $\sim10^{-3}$. In addition to testing the unitarity of the CKM matrix in the charm sector, this ratio is explicitly required to extract the unitarity angle $ \gamma $ from the ratio $B( B_d \rightarrow \rho^0 \psi )/B( B_d \rightarrow K^* \psi )$ discussed earlier. \section{--~ Testing QCD with Charm Hadroproduction} At the parton level, $ c \bar c $ production is supposed to be described by perturbative QCD\cite{NDE}. At the hadron level, the situation becomes more complicated. Several experiments have reported large leading-particle effects at high $x_F$\cite{E769}. Leading-twist factorization in perturbative QCD predicts that the charm quark's fragmentation is independent of the structure of the projectile, while the data indicate that the produced charm quark coalesces or recombines with the projectile spectator. To test models of higher-twist effects\cite{Brodsky}, one wants to look at the observed production asymmetries as functions of $ p_T $ and $ x_F $ jointly. Measuring these asymmetries for different target nuclei ({\it i.e.} measuring the $ A $-dependence of these asymmetries) will provide an extra handle on how quarks evolve into hadrons. \section{--~ Experimental Issues} Building an ultrahigh-statistics charm experiment will be a challenge. The next-generation fixed-target experiments at Fermilab each project reconstructed charm samples of order $ 10^6 $ events. A Tau-Charm Factory operating at a luminosity of $ 10^{33} \,$cm${}^{-2}$\,sec${}^{-1}$, such as that proposed for SLAC\cite{slac343}, would reconstruct about $ 5 \times 10^6 $ charm per year. The $ B $ factories planned for KEK and SLAC will produce of order $ 10^8 $ $ b \overline b $ events per year at design luminosity. However, the number of reconstructed charm will be similar to that projected for the Tau-Charm Factory. HERA-$B$\cite{albrecht94} will produce a sufficient number of $ D $'s in $ p p $ collisions to imagine an ultrahigh-statistics experiment, but the triggering requirements for charm physics differ substantially from those for $ B $ physics, and the data acquisition system is currently designed to operate at 10 Hz. In addition, the current design for the HERA-$B$ vertex detector entails much more multiple scattering and much poorer vertex resolution than are desirable for a charm experiment. There is no clear route to higher luminosities for $ e^+ e^- $ machines or photon beams, so we are left with the problem of working with a relatively small cross-section in a hadronic environment. Whether it is a fixed-target experiment or a collider experiment that we consider, triggering selectively and efficiently will be the first major problem. Building a detector which minimizes backgrounds will be another problem. If we are looking for physics beyond the Standard Model, or looking for relatively rare decays expected within the Standard Model, reducing backgrounds will be at least as important as maintaining high efficiency for the interesting signals. Two examples should suffice: \begin{description} \item [1)~ ] The FCNC decay $ D^+ \rightarrow \pi^+ \mu^+ \mu^- $ is expected to have a branching ratio less than $ 10^{-8} $ in the Standard Model\cite{schwartz93} (except for the decay $ D^+ \rightarrow \phi \pi^+ $ followed by $\phi \rightarrow \mu^+ \mu^- $, which populates a limited region of the Dalitz plot). E791\cite{Nguyen} finds that its sensitivity is greatest when the expected number of background events is between 5 and 10 in the signal region. If one were to scale up from this experiment simply, sensitivity would increase only as the square root of the number of reconstructed charm, since the background would grow linearly with the signal. To increase sensitivity here, it will be important to reduce backgrounds without substantially reducing efficiency for detecting muons. This can be achieved by adding redundancy, {\it e.g.} a second view in the muon detector or a redundant muon-momentum measurement, so that the double muon-misidentification probability becomes approximately the square of the single muon-misidentification probability. \item [2)~ ] To measure the ratio of CKM matrix elements by comparing the decay rates for $ D^+ \rightarrow K^* \ell \nu $ and $ D^+ \rightarrow \rho^0 \ell \nu $, it will be critical to separate pions from kaons with a very high degree of confidence; the reflections of these signals feed into each other directly. A fast RICH technology may suffice, but this is another area where redundancy seems necessary to reduce the confusions which lead to systematic errors. \end{description} Finally, it seems obvious that silicon pixel devices will be necessary to provide both the spatial resolution and the segmentation that are required for unambiguous vertexing in the high-rate small-angle region. \section{--~ Summary} Charm physics provides a window into the Standard Model, and possibly beyond, that complements those provided by other types of experiments. In searches for $D-{\overline D}$ mixing, FCNC, lepton-number-violating decays, or {\it CP}-violating amplitudes, we are probing physics at the TeV level which may not be accessible to other experiments: down-sector and up-sector quarks can couple differently to new physics, and the charm quark is the only up-sector quark for which such studies are possible. Within the Standard Model, charm is probably the best place to test heavy-quark symmetry quantitatively, and it is the best place to measure some of the CKM matrix elements. While ultrahigh-statistics experiments will be extremely difficult, we can reasonably imagine that the technology will exist in the next decade to reconstruct 100 million charm. Getting from here to there will require a substantial R\&D effort, and developing the expertise to design and build such an experiment will require commitment from the individuals who will contribute directly, from the laboratory at which it will be done, and from the community as a whole. We thank the organizers of this Workshop for offering us this opportunity to discuss heavy-quark physics in such attractive surroundings. We also thank our colleagues from Fermilab proposal P829\cite{p829} for their contributions. This work was supported by NSF grant PHY 92-04239 and DOE grant DE-FG02-94ER40840.
\section{Introduction} Perturbation theory has been an important calculational method in quantum field theory for several decades. However the practical usefulness of perturbation theory is limited mainly by the rapid rise in the amount of labour required to improve the order of approximation. For instance, the renormalization group functions can be calculated analytically to three loops in non-abelian gauge theory, which requires the evaluation of 440 three loop diagrams, and to five loops in $\lambda \phi^{4}$ theory~\cite{tar80,che81}. The non-numerical nature of such perturbative calculations in quantum field theory has complicated the effort to automate these calculations. However, the development of specialised computer algebra systems~\cite{str74} and improvements in general purpose computer algebra systems have greatly facilitated this task~\cite{kub90,hy92,ber92,ft92,jam93,west93}. In devising a scheme for the perturbative calculation of amplitudes in quantum field theory the main areas which need to be developed are: \begin{itemize} \item The perturbation expansion itself, \item Lorentz tensor, Dirac and symmetry group algebra, and \item Feynman integration. \end{itemize} The perturbation expansion is either a Dyson-Wick expansion in the canonical formalism or equivalently an expansion of the generating functional in the path integral formalism. The perturbation expansion will not concern us here, nor will we be concerned with automating symmetry group algebra which is simple for all cases of interest to us. Our focus will be on Lorentz tensor and Dirac algebra, and on Feynman integration. There are a number of existing packages that tackle the problem of Dirac algebra computations~\cite{kub90,hy92,ber92,jam93}. The symbolic evaluation of Feynman diagrams at tree level~\cite{ber92} and one loop~\cite{west93} have been approached using Wolfram's computer algebra package Mathematica~\cite{wolf88}. Fleischer and Tarasov~\cite{ft92} present a package for the evaluation of certain two loop Feynman integrals written in the computer algebra language FORM~\cite{ver90}, while symbolic three loop calculations~\cite{lar93} have also been carried using FORM. The current paper presents a package that combines a method for simplifying the Dirac algebra with procedures for evaluating massless scalar or tensor Feynman integrals at one or two loops, and is implemented in Mathematica. The following sections discuss our scheme for representing Feynman diagrams. The scheme is essentially a definition of a notation that can be readily expressed in Mathematica. Section two introduces the notational scheme, section three will discuss some aspects of rules which have been encoded to simplify and integrate expressions, and section four demonstrates the evaluation of some specific diagrams. Usage messages for functions defined by us can be found in the appendix. \section{Notational Scheme} In this section it will be shown how one can use the symbolic algebra capabilities of Mathematica to represent and simplify the various types of tensorial and matrix expressions that arise in amplitudes of gauge theories. This is based upon a suitable notation for dimensionally continued tensors and Dirac gamma matrices. The aim is to establish a notation that can be readily accommodated by the symbolic capabilities of Mathematica. Dirac algebra can be performed without reference to any representation of the Dirac matrices, by using the commutation relations which define the algebra and identities derived from those relations, along with the rules of matrix multiplication. Tensor manipulations are also readily implemented symbolically. As we are interested in evaluating amputated one particle irreducible amplitudes which have been regularised using dimensional regularisation~\cite{tv72}, no explicit representation of Dirac gamma matrices is necessary. The first step is to separate symbols which will represent four-vectors from symbols which will represent tensor indices. We generally use one or more lower case letters and numerals which are not otherwise defined, such as {\tt a,k,p1 \ldots } for four-vectors and indices. The two types of symbols can be distinguished by declaring a list of symbols which will represent four-vectors. Any symbols which appear that are not in this list and are not otherwise defined shall be understood to represent a Lorentz index, and need not be declared. These symbols will appear only as the arguments of Mathematica expressions. The expressions themselves will represent Dirac matrices, Lorentz scalars, four-vectors and tensors depending on the symbols that appear in their arguments. We give the list of declared four vectors the name {\tt momenta}. To illustrate this scheme assume {\tt momenta = \{k,p,q\} } throughout the rest of this section. This list may be enlarged at any time provided symbols that have already been used as Lorentz indices are not included. Let the function {\tt g,} a symmetric function of two arguments, denote the object that will carry the properties of the metric tensor. What {\tt g} actually represents will depend on its arguments as follows: \begin{enumerate} \item {\tt g[a,b]} represents the metric tensor with Lorentz indices {\tt a} and {\tt b,} because {\tt a} and {\tt b} are not in the list {\tt \{k,p,q\},} \item {\tt g[a,k]} represents the four-vector {\tt k} with Lorentz index {\tt a,} because only {\tt k} is in the list {\tt \{k,p,q\},} and \item {\tt g[k,p]} represents the scalar product of the four-vectors {\tt k} and {\tt p} which are both in {\tt \{k,p,q\}.} \end{enumerate} Products of gamma matrices are represented by a function {\tt d} of any non-zero number of arguments. Again, what {\tt d} represents depends on the arguments. Wherever there appears an argument that is a member of {\tt momenta} the corresponding matrix is the contraction of the gamma matrix in that position with the four-vector. Otherwise the matrix is simply a gamma matrix with the symbol denoting a Lorentz index. For example, the object {\tt d[a,b,k,p,b,d,k,q]} denotes \begin{equation} \gamma^{\alpha} \gamma^{\beta} \! \not \! k \not \! p \: \gamma_{\beta} \: \gamma^{\delta} \! \not \! k \not \! q . \end{equation} Free Lorentz indices appearing as arguments of {\tt g} and {\tt d} can represent contravariant or covariant indices. In renormalizing quantum field theories we consider only amputated amplitudes. Information about components of four-vectors and matrices is not required. Thus the distinction between contravariant and covariant components is not important here. One simply takes the rank structure of the final expression to be that of the input expression. If one wishes to consider expressions which included external lines then one would need to distinguish between contravariant and covariant indices. This could be done by declaring all covariant indices in a list which one may call {\tt covariant.} Then symbols which appear in neither of {\tt momenta} or {\tt covariant} and were not otherwise defined would implicitly represent contravariant indices. Traces of products of gamma matrices can be denoted by a function {\tt trace} whose argument is a linear combination of products of gamma matrices. This function will automatically convert the trace of a linear combination to the linear combination of traces before any traces are evaluated. Other tensors may be represented by other functions with the rank structure given implicitly by the arguments. For example one may wish to manipulate the Levi-Civita tensor in four dimensions. One would define a function {\tt e} depending on four arguments. Then, for example {\tt e[a,b,u,v]} would represent \begin{equation} \epsilon^{\alpha \beta \mu \nu} \end{equation} and {\tt e[a,b,k,p]} would represent \begin{equation} \epsilon^{\alpha \beta \mu \nu} k_{\mu} p_{\nu}. \end{equation} The mass of a particle is represented by the function {\tt M[x]} where {\tt x} is a label for the convenience of the user and is not used by the program. Similarly, other scalars are represented by the function {\tt scalar[x].} The left and right helicity projection operators are represented by {\tt L} and {\tt R,} respectively. The matrix $\gamma_{5}$ has not been explicitly represented, nor have traces involving $\gamma_{5}$ been implemented. However the definitions could be easily extended if traces involving $\gamma_{5}$ were required (see for example~\cite{west93}). Manipulation and simplification of expressions can be performed by machine using pattern recognition, procedural programming, rule based programming and functional programming all of which are supported by Mathematica. One can perform basic manipulations such as \begin{itemize} \item contracting repeated indices in tensor expressions, \item simplifying an expression of the form $ \gamma^{\mu} \gamma^{\alpha_{1}} \ldots \gamma^{\alpha_{j}} \gamma_{\mu} $ \item commuting a gamma matrix through one or more gamma matrices, \item the evaluation of the trace of any number of gamma matrices, and \item evaluation of Feynman integrals \end{itemize} and any other lengthy algebraic manipulation that would be prohibitively tedious to perform manually. Some of the rule definitions which are useful for the manipulations described are illustrated in the next section. Note that the system function {\tt Dot} has been used in place of {\tt d} for products of gamma matrices. This function will perform matrix multiplication when explicit matrices in component form are placed in its arguments. Since we require no representation for the gamma matrices and thus only symbols appear as its arguments, the function {\tt Dot} will simply represent non-commutative multiplication and one can assign rules required to invoke the various properties of Dirac algebra which are required. We have chosen to use {\tt Dot} for this purpose because a convenient input notation \begin{equation} {\tt Dot[a1, a2, \ldots] \; \equiv \; a1.a2. \, \ldots} \end{equation} is available. Since {\tt Dot} has the attribute {\tt OneIdentity} the expression {\tt Dot[a]} is equivalent to {\tt a} where {\tt Head[a] = Symbol.} Hence a solitary symbol which is otherwise undefined represents a single gamma matrix. Together with symbols {\tt i} for the imaginary unit, {\tt n} for the number of space-time dimensions, and {\tt eps = 4-n,} this scheme is sufficient to represent any amputated diagram in a gauge theory. In the next section we will discuss how to manipulate, simplify, and integrate expressions within this scheme. \section{Implementation of Symbolic Algebra} We will begin by discussing expressions representing Lorentz tensors. This will be followed with Dirac algebra and expressions involving both gamma matrices and tensors, and traces of gamma matrices. Finally the evaluation of two-loop Feynman integrals will be discussed. In evaluating Lagrangian counterterms, which in turn give renormalization group functions, we are interested in the pole part of amputated diagrams. This necessitates the evaluation of Laurent expansions and some points relating to this will be discussed briefly. The metric of $n$-dimensional Minkowski space is represented by the symmetric function {\tt g} with two arguments, \begin{eqnarray} \lefteqn{{\tt Attributes[g] \,=\, \{Orderless\} }} & & \\ \lefteqn{{\tt g[x\_,x\_] \, := \, n}} & {\tt /; \,FreeQ[momenta,x]} & \\ \lefteqn{{\tt g[a\_,b\_]\hat{\:}2 \, := \, n}} & & \nonumber \\ & \hspace{1.4cm}{\tt /; FreeQ[momenta,a] \: \&\& \: FreeQ[momenta,b]} & \\ \lefteqn{{\tt g[a\_,b\_]\hat{\:}2 \, := \, g[b,b]}} & & \nonumber \\ & \hspace{1.4cm}{\tt /; FreeQ[momenta,a] \: \&\& \: MemberQ[momenta,b] } & \hspace{2.0cm} \end{eqnarray} Note that the action of the rules is conditional upon whether or not one or both arguments are in the list {\tt momenta.} In conventional notation the rules stated are \begin{eqnarray} && g^{\alpha}_{\alpha} = n \\ && g^{\alpha \beta} g_{\alpha \beta} = n \\ && b^{\alpha}b_{\alpha} = b^{2} \end{eqnarray} where b is a vector. Rules can be applied at the discretion of the user by defining new functions which act on tensors. For example the relations \begin{eqnarray} && g^{\alpha \beta} g_{\beta \lambda} = \delta^{\alpha}_{\lambda} \label{eq:rmind} \\ && b^{\alpha} g_{\alpha \lambda} = b_{\lambda} \label{eq:lower} \end{eqnarray} can be implemented by \begin{eqnarray} \lefteqn{ {\tt cc[x\_,a\_,b\_,s\_] \: := \: x \, /. \,g[a,s]\, g[s,b]\: \rightarrow \: g[a,b]} } \nonumber \\ && \hspace{6.0cm} /; \, {\tt FreeQ[momenta,s]} \hspace{3.0cm} \end{eqnarray} Note that in this notation both (\ref{eq:rmind}) and (\ref{eq:lower}) are embodied in one rule. Functions can appear in the definitions of functions. For instance, a function which contracts all occurrences of a repeated index is valuable. Suppose all the arguments of {\tt g} in an expression {\tt x} are placed in a list called {\tt args} except for one repeated index {\tt s.} We have defined a function, {\tt contract[x, args, s],} which will perform all contractions of the repeated index {\tt s} in {\tt x.} The definition of contract involves two other functions, {\tt cc} which controls the actual replacement that is made, and {\tt ll} which takes a list of symbols and returns a list of all pairs of those symbols. Extending this further, we have defined a function, {\tt contractall[x, repeated, other],} which takes an expression {\tt x,} a list of all repeated indices placed in the second argument, and a list of all remaining arguments of {\tt g} in {\tt x} in the third argument and performs the contraction of all repeated indices. The definition of this function is in terms of {\tt contract.} As for {\tt g}, rules are assigned to {\tt Dot} to automatically implement properties of matrix algebra. Note that since {\tt Dot} is a system function it must be unprotected before rules can be added to its definition. The rules we define for {\tt Dot} are applied automatically to any input expression until no further changes occur. These rules correspond to properties such as the distributive law over addition and scalar multiplication. We have chosen to use the symbol {\tt J} to denote the identity matrix. Expressions which represent a combination of tensors and gamma matrices will involve both {\tt g} and {\tt Dot.} In cases where a Lorentz index appears in an argument of both {\tt g} and {\tt Dot} the index may be contracted out of the expression. A function {\tt slash} is defined in the appendix which performs this task. Identities, such as commutation relations can be applied at the discretion of the user by defining functions which take an expression, search for a specified pattern and replace the pattern by an equivalent expression. For example, the commutation relations \begin{eqnarray} && \gamma_{\alpha} \gamma_{\beta} = - \gamma_{\beta} \gamma_{\alpha} + 2 g_{\alpha \beta} \\ && \gamma_{\alpha} \gamma_{\beta} \gamma_{\lambda} = \gamma_{\beta} \gamma_{\lambda} \gamma_{\alpha} + 2 \gamma_{\lambda} g_{\alpha \beta} - 2 \gamma_{\beta} g_{\alpha \lambda} \end{eqnarray} are applied by the function {\tt comm}, defined as \begin{eqnarray} \lefteqn{ {\tt comm[x\_,a\_,b\_] \: := \: x \, /. \,a.b \rightarrow b.a \: + \: 2J \, g[a,b]}} & & \\ \lefteqn{{\tt comm[x\_,a\_,b\_,c\_] \: := \: x \, /. \, a.b.c \rightarrow b.c.a \: + \: 2c\, g[a,b] \: - \: 2b\, g[a,c]}} & & \hspace{12.0cm} \end{eqnarray} Which of the two rules is applied depends on how many arguments are passed to {\tt comm}. Note that the rules are valid whether or not any of {\tt a, b} and {\tt c} appear in the list {\tt momenta.} Further rules have been defined for commuting through more than two gamma matrices, and another function called {\tt rcomm} for commuting in the reverse order. Rules which apply identities for expressions of the form \begin{eqnarray} &&\gamma_{\mu} \gamma_{\alpha_{1}} \ldots \gamma_{\alpha_{j}} \gamma^{\mu} \\ && k\!\!\!/ \gamma_{\alpha_{1}} \ldots \gamma_{\alpha_{j}} k\!\!\!/ \; , \;j=1,2,3,\ldots \end{eqnarray} have been similarly defined. For instance \begin{eqnarray} {\tt con[x\_,a\_] \: := } & {\tt If \left[ \:MemberQ[momenta,a] \, , \right. }\hspace{5.5cm} & \nonumber \\ & {\tt x\, /. \, a.a \rightarrow g[a,a]\, J,}\hspace{2.0cm} & \nonumber \\ & {\tt x\, /. \, a.a \rightarrow n\, J }\hspace{3.0cm} & \nonumber \\ & \lefteqn{\left. \right]} \hspace{8.5cm} & \end{eqnarray} \begin{eqnarray} {\tt con[x\_,a\_,b\_]\: := } & {\tt If \left[ MemberQ[momenta,a] \, , \right. } \hspace{5.5cm} & \nonumber \\ & {\tt x\, /. \, a.b.a \rightarrow - b\, g[a,a]\: +\: 2a\, g[a,b],} & \nonumber \\ & {\tt x\, /. \, a.b.a \rightarrow (2-n)\, b}\hspace{2.5cm} & \nonumber \\ & \lefteqn{\left. \right]} \hspace{8.5cm} & \end{eqnarray} The answer is dependent on whether or not the repeated symbol {\tt a} appears in the list {\tt momenta.} Rules for $j>1$ can be included as needed. To simplify expressions involving helicity projection operators, the functions {\tt movel} and {\tt mover} have been defined. They, respectively, move the helicity projection operators to the left or to the right in each term in an expression. The trace of an arbitrary linear combination of products of gamma matrices can be evaluated using the linearity property of traces and the recursive relation \begin{eqnarray} Tr(\gamma^{\alpha_{1}} \gamma^{\alpha_{2}}) & = & 4 g^{\alpha_{1} \alpha_{2}}\\ Tr(\gamma^{\alpha_{1}} \ldots \gamma^{\alpha_{2j}}) & = & \sum_{l=o}^{2j-2} (-1)^{l} g^{\alpha_{1} \alpha_{2j-l}} Tr(\gamma^{\alpha_{2}} \ldots \hat{\gamma}^{\alpha_{2j-l}} \ldots \gamma^{\alpha_{2j}}) \label{eq:tr} \\ Tr(\gamma^{\alpha_{1}} \ldots \gamma^{\alpha_{2j+1}}) & = & 0 \end{eqnarray} where the hat over the gamma matrix denotes its absence~\cite{west93}. The function {\tt trace} will firstly convert the trace of a linear combination into a linear combination of traces, and then applies a function {\tt tr} which makes use of the recursion relation (\ref{eq:tr}) to evaluate the traces. Note that {\tt tr} first checks that the length of the product is even and returns zero if it is not. The case of a product of two gamma matrices acts as an initial condition. The scheme described so far can be used to represent any Feynman integral which can arise in a gauge theory. A means of evaluating the integrals would complete a scheme for the evaluation of amputated Feynman diagrams. We will present a simple procedure which can be adapted to all cases. We begin by considering integrals of the form \begin{equation} \mu^{\lambda} \int \frac{d^{n}q}{(2 \pi)^{n}} \frac{q_{\alpha} q_{\beta} \ldots}{(q^{2})^{j} ((q \pm p)^{2})^{l}} \label{eq:stdint1} \end{equation} and \begin{equation} \mu^{\lambda} \int \frac{d^{n}q}{(2 \pi)^{n}} \frac{q \!\!\!/ q_{\alpha} q_{\beta} \ldots}{(q^{2})^{j} ((q \pm p)^{2})^{l}} \label{eq:stdint2} \end{equation} where $\mu$ is an arbitrary mass scale, $q$ is the integration variable and the rank of the tensor in the numerator is usually less than or equal to 3. The factor $\mu^{\lambda}$ is introduced to ensure that the integral is dimensionless for any $n$ (hence $\lambda$ depends on $n$). This is sufficient for most two loop calculations but if rules for the integration of higher rank expressions are needed, then extra rules can be included. The main consideration in defining an integration function is to be certain that there is no residual dependence on the integration variable in the final expression. For example, a function could be defined which searches for the pattern \begin{equation} {\tt \frac{1}{g[q,q] \: g[q+p,q+p]}} \end{equation} and replaces it using a replacement rule. If that function were then applied to a pattern like \begin{equation} {\tt \frac{T[q]}{g[q,q] \:g[q+p,q+p]}} \end{equation} the integration function would treat this as a product of two factors \begin{equation} {\tt T[q]\; \frac{1}{g[q,q] \:g[q+p,q+p]}} \end{equation} and replace only the second factor, leaving {\tt T[q]} in the resulting expression. Hence it is important to check the final expression for the presence of the integration variable. In the integrals (\ref{eq:stdint1}) and (\ref{eq:stdint2}) there are other considerations. The rank, the symbols used for indices, the integration variable, the parameter {\tt p,} and the powers {\tt j} and {\tt l} should all be variables which can be passed to the integrating function. We have defined functions {\tt inttensor} and {\tt intgamma} which will return the integrals of (\ref{eq:stdint1}) and (\ref{eq:stdint2}) respectively. The arguments to be passed to {\tt inttensor} and {\tt intgamma} are \begin{equation} {\tt inttensor[x, q, p, a, b, \ldots, j, l] } \end{equation} and \begin{equation} {\tt intgamma[x, q, p, a, b, \ldots, j, l], } \end{equation} where {\tt x} is the expression to be integrated, {\tt q} is the integration variable, {\tt p} is the four-momentum parameter, {\tt a,b,\ldots} are Lorentz indices, and {\tt j} and {\tt l} are the powers in the denominator. Again the definitions of {\tt inttensor} and {\tt intgamma} consist of a limited series of rules. Which of the rules is applied depends on the number of arguments passed to {\tt inttensor} and {\tt intgamma.} If the rules do not cover cases of high rank that may be needed then new rules may be included as needed. Rank three integrals are usually sufficient for most two-loop calculations. Integrals involving non-zero masses have not been implemented, though a function for these could be defined. Dimensionally regularised integrals can be expanded as a Laurent series in $\epsilon \; = \; 4-n$. In general an integral over $l$ four vectors will have a leading pole of order $\epsilon^{-l}$. To evaluate counterterms only pole parts of the Laurent expansions of integrals need be evaluated. A table of pole parts of integrals can be prepared with the help of {\tt inttensor} and {\tt intgamma.} This table can then be used to evaluate the pole parts of diagrams and hence the counterterms. The integrals required for this can be classified according to the form of the denominators of the Feynman integrals. For example, to calculate the two-loop boson self energy in a $SU(m)$ gauge theory with massless fermions there is just one class of integrals as the Feynman integral is always of the form \begin{equation} \mu^{\lambda} \int_{q,k} \frac{T_{\mu \nu}(q,k,p)}{q^{2}(q+p)^{2}k^{2}(k+p)^{2}(k-q)^{2}} \label{eq;twopt} \end{equation} where $p$ is the external momentum, and $T_{\mu \nu}(q,k,p)$ is a rank-2, dimension 4 tensor constructed from $q$, $k$, and $p$. There are 75 distinct possibilities for $T_{\mu \nu}$ but fortunately many of the corresponding integrals are related, and the number of independent integrals is reduced to about 10. If we are interested only in the pole parts then we can express any integral of the form (\ref{eq;twopt}) in terms of the pole parts of just four integrals. We choose the four integrals where $T_{\mu \nu} (q,k,p)$ is either \begin{equation} t_{\mu \nu}(p) (q+p)^{2} \end{equation} or \begin{equation} t_{\mu \nu}(p) (k-q)^{2} \end{equation} and $t_{\mu \nu}(p)$ is either $p_{\mu} p_{\nu}$ or $g_{\mu \nu} p^{2}$. In the integrals discussed we have represented the Euler gamma function by {\tt gam[x].} For the evaluation of pole parts of one-particle irreducible amplitudes one must be able to perform the Laurent expansions. This can be done by either replacing {\tt gam} with the system defined version of the Euler gamma function and using such functions as {\tt Series,} or by defining a sufficient set of rules for {\tt gam} to evaluate the expansion. \section{Examples} As an illustration of the application of this work to some specific examples, we will briefly discuss the evaluation of the integral \begin{equation} \mu^{2(4-n)} \int \frac{d^{n}q}{(2 \pi)^{n}} \frac{d^{n}k}{(2 \pi)^{n}} \frac{1}{k^{2} (k+p)^{2} q^{2} (q-k)^{2}} \label{eq;intex} \end{equation} the evaluation of the diagram shown in figure~\ref{fig;bosonse}, and the simplification of the diagram in figure~\ref{fig;quarkem}. Diagram~\ref{fig;bosonse} is a contribution to the gauge boson self energy in the presence of massless fermions in some representation R of the gauge group. The diagram~\ref{fig;quarkem} represents a contribution to the quark-photon vertex in the Weinberg-Salam model. Before evaluating the integral (\ref{eq;intex}) we must first declare the list of four-vector symbols as \begin{equation} {\tt momenta = \{k, p, q, k+p, q+p, k-q, q-k\} \label{eq;mlist}} \end{equation} Then we specify the integrand by \begin{equation} {\tt x = \frac{1}{g[k, k]\, g[k+p, k+p]\, g[q, q]\, g[q-k, q-k]}} \end{equation} The integral over {\tt q} is performed using inttensor, \begin{equation} {\tt x= inttensor[x, q, k, 1, 1] } \end{equation} which results in an expression depending on the integration variable {\tt k.} The integrand of this expression is proportional to \begin{equation} {\tt \frac{1}{(g[k,k])^{1+ \frac{eps}{2}} \, g[k+p, k+p]}} \end{equation} and can also integrated by using {\tt inttensor} as follows. \begin{equation} {\tt x= inttensor[x, k, p, 1+eps/2, 1]} \end{equation} The resulting expression is \begin{equation} {\tt \frac{-(2^{2 eps} \, mu^{2 eps}\, pi^{-4\,+\,eps}\, gam[1-eps]\, gam[1- \frac{eps}{2}]\, gam[\frac{eps}{2}]\, gam[eps])} {256\, (-1)^{eps}\, g[p,p]^{eps}\, gam[2 -\frac{3 eps}{2}]\, gam[2-eps]\, gam[1+\frac{eps}{2}]}} \end{equation} where {\tt mu} is the introduced mass scale $\mu$, and this can be Laurent expanded to order $\epsilon^{-1}$. We obtain \begin{equation} {\tt \frac{-2 \: + \: 2 \, elog \: - \: 5 \, eps}{256\, eps^{2}\, pi^{4}} + O( eps)^{0}} \end{equation} where \begin{equation} {\tt elog = egam - log \left( \frac{4 \, pi \, mu^{2}}{-g[p,p]} \right)} \end{equation} and {\tt egam = 0.577\ldots} is Euler's constant. This gives the result for two of the four integrals required to determine the pole part of (\ref{eq;twopt}) and the remaining integrals can be evaluated similarly. Then to integrate the expression (\ref{eq;twopt}) we first replace the integrand with an expression that has the same pole part and then replace the integrals with the Laurent expansions. The functions we have defined to do this are {\tt samepole} and {\tt poleform} respectively. To evaluate the diagram of figure~\ref{fig;bosonse} we declare the momenta list to be~(\ref{eq;mlist}), and define some initial expressions. In the notation of our scheme they are the coefficient of the integral, the numerator and denominator of the integrand. The coefficient is \begin{equation} {\tt coefficient = \frac{i}{2} \, g^{4} \,C[G] \, T[R] \, delta[a, b]} \end{equation} where {\tt a} and {\tt b} are gauge group indices, {\tt delta} is the delta symbol and {\tt C[G]} and {\tt T[R]} are gauge group factors. The numerator of the integrand has two factors \begin{eqnarray} {\tt y = (g[p,a]\, -\, g[k,a]) \, g[u,l] \: +\: (2 g[k,u]\, +\, g[p,u])\, g[l,a] -} & & \nonumber \\ {\tt (2 g[p,l] \, + \, g[k,l])\, g[a,u]}\hspace{4.0cm} & & \end{eqnarray} and \begin{equation} {\tt z = trace[a.(p+q).v.q.l.(k-q)]} \end{equation} and the denominator is \begin{equation} {\tt denominator = g[q,q]\, g[q+p,q+p]\, g[k,k]\, g[k+p,k+p]\, g[k-q,k-q]} \end{equation} In the numerator there are two repeated indices. We set ${\tt repeated = \{a,l\},}$ ${\tt other=\{u,v,k,p,q\} }$ and ${\tt x = Expand[y\: z]}$ and remove the repeated indices from {\tt x} using {\tt contractall,} \begin{equation} {\tt x=contractall[x,repeated,other] } \end{equation} All scalar products are then eliminated using {\tt preps} and {\tt prepd} which make use of identities such as \begin{equation} k.p = \frac{1}{2} ((k+p)^{2} -k^{2} -p^{2}). \end{equation} The resulting expression for {\tt x} consists of 69 terms. We then replace {\tt x} with an expression with the same pole part using {\tt samepole.} The result is divided by {\tt denominator} and the Laurent expansion is evaluated by applying {\tt poleform.} Simplifying the result and multiplying by {\tt coefficient} gives the final result for the diagram \begin{eqnarray} & & {\tt -( g^{4}\, i\, C[G]\, T[R]\, TR[J]\, delta[a,b] } \nonumber \\ & & \hspace{0.4cm} {\tt ((-48\: -\: 76 eps\:+\:48 \, elog \: eps)\, g[p,u] \, g[p,v] \:+ } \nonumber \\ & & \hspace{0.4cm} {\tt (12\: -59\, eps\: -\: 12 \, elog\: eps)\, g[p,p]\, g[u,v]))/(55296 \, eps^2 \, pi^4) } \end{eqnarray} where {\tt TR[J]} is the trace of the spin identity matrix. We have used our scheme to evaluate the gauge boson, fermion and ghost anomalous dimensions to two loops. We have also calculated the coupling constant renormalization by evaluating the gauge boson - ghost vertex, and evaluated the Callan-Symanzik beta function. Our results are in complete agreement with~\cite{ego79} for these calculations. Now consider the diagram in figure~\ref{fig;quarkem}, which is a two loop diagram contributing to the quark-photon vertex. The one loop, flavour-changing quark self energy, with momentum $p$ can be written as \begin{equation} A \, p\!\!\!/ L + B\, p\!\!\!/ R + C\, L + D \, R \end{equation} where $A, B, C$ and $D$ are scalars depending upon $p^{2}$ and the quark masses. After defining \begin{equation} {\tt momenta = \{k,p\} } \end{equation} the numerator of the diagram in figure~\ref{fig;quarkem} is proportional, in our notation, to \begin{eqnarray} {\tt x} & = & {\tt alpha \, .\, L\, .\, (p+k+M[s])\, .\, (scalar[A]\, (p\, +\, k)\, .\, L + scalar[B]\, (p\, +\, k)\, .\, R } \nonumber \\ & & \hspace{0.4cm} {\tt + scalar[C]\, L + scalar[D]\, R)\, .\, (p+k+M[d])\, .\, mu \, .\, (p-k+M[d])\, .\,} \nonumber \\ & & \hspace{0.8cm}{\tt alpha. L } \end{eqnarray} The masses are implicitly multiplied by the identity matrix. To simplify this we first apply \begin{equation} {\tt x = movel[x]} \end{equation} Then repeated application of the function {\tt con} will do the contractions over $\gamma_{\alpha}$ and contract the pairs of $k\!\!\!/$'s and $p\!\!\!/$'s. For example, the following does all possible contractions over $k\!\!\!/$ and $p\!\!\!/$. \begin{eqnarray} & & {\tt x\, =\, con[x,p]\, ;\, x\, =\, con[x, k]\, ;\, x\, =\, con[x, p,mu]\, ; } \nonumber \\ & & \hspace{0.4cm} {\tt \, x\, =\, con[x,k,mu]\, ;\, x\, =\, con[x,p,k]\, ;\, x\, =\, con[x,k,p]\, ; } \nonumber \\ & & \hspace{0.8cm} {\tt x\, =\, con[x,p,mu]\, ;\, x\, =\, con[x,k,mu] } \end{eqnarray} A further seven applications of {\tt con} are necessary to do all the contractions over $\gamma_{\alpha}$. In the resulting expression, repeated use of {\tt comm} \begin{eqnarray} & & {\tt x\, =\, comm[x,p,k]\, ; \, x\, =\, comm[x,p,mu]\, ;\, x\, =\, comm[x,k,mu]\, ; } \nonumber \\ & & \hspace{0.4cm} {\tt x\; =\; comm[x,p,k] } \end{eqnarray} followed by collecting with respect to {\tt Dot,} produces a final expression of four terms, proportional to $R \gamma_{\mu} k\!\!\!/ p\!\!\!/$, $R \gamma_{\mu}$, $R k\!\!\!/$ and $R p\!\!\!/$. The full expression has seventy-six terms when expanded and is too long to reproduce here. \section{Conclusion} We have presented a scheme for representing amputated Feynman diagrams beyond tree level in terms of Mathematica expressions and have shown how the algebra involved in simplifying these expressions and evaluating the integrals can be automated using the symbolic algebra capacity of Mathematica. This scheme is based upon the specification of a sufficiently convenient notation for Lorentz tensors and Dirac algebra. An automatic procedure for the evaluation of pole parts for dimensionally regularised massless Feynman integrals has been used to evaluate two loop counterterms in a non-abelian gauge theory with fermions and for the simplification of amputated amplitudes in the Weinberg-Salam model. The methods used in this paper can be readily extended to handle integrals of higher loop order, and integrals involving masses. Our approach succeeds in eliminating all tedious hand calculations. The time required to perform calculations automatically is generally small compared to the time required to prepare input and organise a calculational sequence. \section{Acknowledgements} We would like to thank A. A. Rawlinson and A. C. Kallionatis for their useful suggestions and assistance. \section{Appendix} This appendix contains the usage messages for functions described in this article. The following usage messages are for functions which simplify products of Dirac gamma matrices. \begin{verse} {\tt comm::usage = "comm[x, a, b1, b2, \ldots] commutes a to the} \\ {\tt right through b1.b2 \ldots wherever the pattern a.b1.b2 \ldots } \\ {\tt appears in x (up to 5 b's programmed)."} \end{verse} \begin{verse} {\tt rcomm::usage = "rcomm[x, b1, b2, \ldots, a] commutes a to} \\ {\tt the left through b1.b2 \ldots wherever the pattern b1.b2 \ldots .a} \\ {\tt appears in x (up to 5 b's programmed)."} \end{verse} \begin{verse} {\tt con::usage = "con[x, a, b1, b2, \ldots] evalates a.b1.b2 \ldots .a} \\ {\tt by contracting out the a's. Replacement made wherever the} \\ {\tt pattern appears in x (Up to 6 b's programmed)."} \end{verse} \begin{verse} {\tt trace::usage = "trace[x] evaluates the trace of a linear} \\ {\tt combination of products of gamma matrices that does not} \\ {\tt include gamma\_5."} \end{verse} \begin{verse} {\tt movel::usage = "movel[x] commutes all helicity projection} \\ {\tt operators to the left most position in each term in the} \\ {\tt expression x."} \end{verse} \begin{verse} {\tt mover::usage = "mover[x] commutes all helicity projection} \\ {\tt operators to the right most position in each term in the} \\ {\tt expression x."} \end{verse} The following usage messages are for functions which simplify Lorentz tensors. \begin{verse} {\tt contract::usage = "contract[x, other, s] will eliminate} \\ {\tt repeated index s from x. A list other, of all other} \\ {\tt symbols appearing in the arguments of g's must be passed} \\ {\tt to the function."} \end{verse} \begin{verse} {\tt contractall::usage = "Given a list of repeated indices,} \\ {\tt repeated, and a list, other, of any indices appearing in} \\ {\tt the arguments of g's that are not in repeated,} \\ {\tt contractall[x, repeated,other] will eliminate all} \\ {\tt of the repeated indices from x."} \end{verse} \begin{verse} {\tt slash::usage = "slash[x, a, b] finds all patterns of the} \\ {\tt form c1.c2. \ldots .ci.a.d1.d2. \ldots .dj*g[a,b] in x and} \\ {\tt replaces them with c1.c2. \ldots .ci.b.d1.d2. \ldots .dj} \\ {\tt where either or both of i and j may be zero."} \end{verse} \begin{verse} {\tt preps::usage = "preps[x, k, p] replaces g[k, p] by} \\ {\tt (g[k+p, k+p] - g[k, k] - g[p, p])/2 everywhere in x."} \end{verse} \begin{verse} {\tt prepd::usage = "prepd[x, k, p] replaces g[k, p] by} \\ {\tt (g[p, p] + g[k, k] -g[k-p, k-p])/2 everywhere in x."} \end{verse} The following usage messages are for functions which perform Feynman integration. \begin{verse} {\tt inttensor::usage = "inttensor[x, q, p, a, b, \ldots , j, l]} \\ {\tt integrates tensors (g[q, a] g[q, b] \ldots )/(g[q, q]$\hat{\:}$j g[q $\pm$} \\ {\tt p, q $\pm$ p]$\hat{\:}$l) of low rank, where the integration variable} \\ {\tt q, external momentum p, powers j and l, and indices} \\ {\tt a,b \ldots are specified in the arguments of inttensor."} \end{verse} \begin{verse} {\tt intgamma::usage = "intgamma[x, q, p, a, b, \ldots , j, l]} \\ {\tt integrates matrices (q\, g[q, a]\, g[q, b] \ldots )/(g[q, q]$\hat{\:}$j g[q} \\ {\tt $\pm$ p, q $\pm$ p]$\hat{\:}$l) of low rank, where the integration} \\ {\tt variable q, external momentum p, powers j and l, and} \\ {\tt indices a,b \ldots are specified in the arguments of intgamma."} \end{verse} \begin{verse} {\tt samepole::usage = "samepole[x, p] replaces x, the numerator} \\ {\tt of a Feynman integral for a gauge boson self energy diagram} \\ {\tt with a simpler expression that has the same pole} \\ {\tt part (p is the external momentum)."} \end{verse} \begin{verse} {\tt poleform::usage = "poleform[x, p] evaluates the pole part} \\ {\tt of x, a Feynman integral for a gauge boson self energy} \\ {\tt diagram, after samepole has been applied to the numerator."} \end{verse} \pagebreak
\section{Introduction} In a magnetic field the electronic density is elongated along the field direction. There are many intuitive ways of seeing this. For example, in a path integral description, since the zero field propagator is multiplied by the phase factor $exp \bigl[ {ie \over {{\hbar}c}}\oint{\bf A}\cdot d\bf l\bigr]$, each closed path is reweighted by a phase proportional to the magnetic flux it encloses. In the average over paths which produces the electronic density, paths parallel to the field, enclosing no flux, receive a weight of one; while paths perpendicular to the field, enclosing flux, receive a reduced weight. The resulting average for the electronic density, thus, has an elongated shape.\ The lowest order statistical model however leads to an isotropic density[1]. The anisotropy requires gradient corrections [2]. The necessity of going beyond the lowest order has also spawned a variety of approximate methods to regain anisotropy [3,4,5,6,7,8].\ In this Letter, we first review the fundamental results of CDFT. Then, neglecting the current variation, we rederive the gradient corrections of Tomishima and Shinjo. Finally, taking this variation into account, we calculate all second order electronic density gradient corrections. \section{Current density functional formalism} For systems in a magnetic field, Vignale and Rasolt [9] show that the density $\rho$ and the paramagnetic current density ${\bf j}_p$, related to the gauge-invariant total current density, by \begin{eqnarray} {\bf j({\bf r})} = {\bf j_p({\bf r})} +{e\over mc}\rho({\bf r})A({\bf r}) \label{current} \end{eqnarray} uniquely determine the external scalar, $V_{x}({\bf r})$, and vector potential $\bf A({\bf r})$. The ground state energy functional, which may be written as \begin{eqnarray} & &E[\rho,{\bf j}_p] = K[\rho,{\bf j}_p] +\int d{\bf r} \, \rho({\bf r}) V_{x} ({\bf r}) + {e^2\over2} \int \int d{\bf r}\, d{\bf r}' {{\rho({\bf r})\rho({\bf r}')}\over{|{\bf r}-{\bf r}'|}} \nonumber\\ & & \mbox{\ \ \ \ \ \ \ \ \ \ \ \ \ } + {e \over c}\int d{\bf r} \, {\bf j}_p({\bf r}){\bf A}({\bf r}) + {e^2\over 2mc^2}\int d{\bf r} \, \rho({\bf r}){\bf A^2({\bf r})} + E_{xc}[\rho,{\bf j}_p] \label{energie} \end{eqnarray} where $K[\rho,{\bf j}_p]$ is the kinetic energy of a free electron gas in the external potentials $V_{x}({\bf r})$ and $\bf A({\bf r})$, and $E_{xc}[\rho,{\bf j}_p]$ is the exchange-correlation energy, is minimized at the ground state energy by the correct density and current distribution. In the spirit of Thomas-Fermi approximations we subsequently ignore the exchange-correlation term and the Hartree mean field term. Although $K[\rho,{\bf j}_p]$ is not invariant under a gauge transformation: \begin{eqnarray} & &\bf A({\bf r})\to \bf A({\bf r})-\bf \nabla\chi \nonumber\\ & &\bf j_p({\bf r})\to \bf j_p({\bf r})+ {e\over mc}\rho({\bf r}) \bf \nabla\chi \label{gauge} \end{eqnarray} the difference: \begin{eqnarray} U[\rho,{\bf j}_p] = K[\rho,{\bf j}_p]- {m\over2}\int d{\bf r} \, {{\bf j_p\bf j_p}\over\rho}=E[\rho,{\bf j}_p]-\int d{\bf r} \, \rho({\bf r}) V_{x}({\bf r}) -{m\over2}\int d{\bf r} \, {{\bf j\bf j}\over\rho} \label{Uen} \end{eqnarray} which depends on the energy and physical current, is gauge invariant and, since unchanged by the addition of the gradient of an arbitrary function, depends on ${\bf j}_p$ only through \mbox{$\bf \nu({\bf r})=\bf \nabla\times \bigl(\bf j_p /\rho\bigr)$}, the vorticity introduced by Vignale and Rasolt [9]. In the next two sections we determine the gradient expansion for $U$. \section{Gradient corrections in strong fields} We consider an electron gas in an uniform magnetic field ${\bf B}= B_{0}\hat{{\bf z}}$. As a first step we will neglect the variation of vorticity. Then the functional U is reduced to a functional of only one variable, $\rho$. We postulate an expression for U with gradient terms: \begin{eqnarray} U[\rho] = \int d{\bf r} \, \biggl\{ f(\rho({\bf r}))+g_{\bot}(\rho({\bf r})){\bigl|} {\bf \nabla}_{\bot}\rho {\bigr|}^2+g_{\|}(\rho({\bf r})){\bigl|} {\bf \nabla}_{\|}\rho {\bigr|}^2 \biggr\} \label{U} \end{eqnarray} where $\bot$ and $\|$ denote orientations perpendicular and parallel to the field. The function $f(\rho({\bf r}))$ can be determined from the Euler equation \begin{eqnarray} \begin{array}{c}\underline{\delta E[\rho({\bf r})]}\\\delta\rho({\bf r}) \end{array}=\mu\;. \end{eqnarray} For the assumed $U$ this gives \begin{eqnarray} & &f'(\rho({\bf r})) -2g_{\bot}(\rho({\bf r})) {\bf \nabla}_{\bot}^2\rho -2g_{\|}(\rho({\bf r})) {\bf \nabla}_{\|}^2\rho \nonumber\\ & & \mbox{\ \ \ \ } - g_{\bot}'(\rho({\bf r})){\bigl|} {\bf \nabla}_{\bot}\rho {\bigr|}^2 -g_{\|}'(\rho({\bf r})){\bigl|} {\bf \nabla}_{\|}\rho {\bigr|}^2=-V_{x}({\bf r})+\mu \label{equa} \end{eqnarray} which, at uniform density $\rho_0$ and no external potential, is simplified to: \begin{eqnarray} f'[\rho_0]=\mu_0 \label{equa1} \end{eqnarray} where $\mu_0$ is the chemical potential of a free electron gas. $g_{\bot}(\rho({\bf r}))$ and $g_{\|}(\rho({\bf r}))$ will be determined by comparing the calculated linear density response to a variation in the external potential, $\delta\rho(\bf q)=\chi(\bf q)\delta{V_x}(\bf q)$, to the free electron gas results. A second variation of the Euler equation gives (after a Fourier transform), \begin{eqnarray} f''(\rho_0) +2g_{\bot}(\rho_0){\bf k}_{\bot}^2 +2g_{\|}(\rho_0){k}_{\|}^2=-1/\chi(\bf k) \label{equa3} \end{eqnarray} at uniform density $\rho_0$. As a preliminary example recall the case of no magnetic field. The chemical potential $\mu=\hbar^{2}k_{F}^{2}/2m= \hbar^{2}(3\pi^{2}\rho)^{2/3}/2m$ so eqn.~(\ref{equa1}) reproduces the usual Thomas-Fermi result, \begin{eqnarray} f(\rho)={3\over 10}{\hbar^{2}\over m}(3\pi^{2})^{2/3}\rho^{5/3} \end{eqnarray} and the Lindhard response formula \begin{eqnarray} \chi( k)=-{m k_{F}\over \hbar^{2} \pi^{2}}\left[ {1\over 2}+ \begin{array}{c}\underline{ 2( k_{F}^{2}- k^{2})}\\ k_{F} k \end{array} \ln\left| \begin{array}{c}\underline{2 k_{F}- k}\\2 k_{F}+ k\end{array} \right| \right] =-{m k_{F}\over \hbar^{2} \pi^{2}}\left(1-{1\over 3}( k/2 k_{F})^{2}\right) +O(k^{4}) \end{eqnarray} or \begin{eqnarray} \chi^{-1}( k)=-{\hbar^{2}\pi^{2}\over m k_{F}}\left(1+{1\over 3}( k/ 2 k_{F})^{2} \right)+ O( k^{4})\;. \end{eqnarray} gives, on comparison with eqn.~(\ref{equa3}), the von Weizsacker gradient corrections \begin{eqnarray} g_{\bot}=g_{\|}={\hbar^{2} \over 3 m}{\pi^{2}\over 8 k_{F}^{3}}= {1\over 9}{1\over 8 \rho}{\hbar^{2}\over m} \end{eqnarray} (with the $1/9$ factor from the long wavelength comparison)[10]. In the case of very strong fields, where all the electrons are in the lowest Landau level and spin polarized, the free electron gas chemical potential $\mu_0=\hbar\omega_{c}/2+ \hbar^{2} k_{F}^{2}/2m$ with $k_{F}=2\pi^{2}\hbar\rho/m\omega_{c}$ so \begin{eqnarray} f(\rho)={{\hbar\omega_c} \over 2}\rho+{{2\pi^4\hbar^4}\over {3m^3{\omega_c}^2}}{\rho}^3 \label{equa2} \end{eqnarray} The free electron gas susceptibility for this case is well known since Horing [11,12] \begin{eqnarray} \chi({\bf k})=-{2\over {4{\pi^2}{\omega_c}l^4 \hbar}} {{F({\bf k}_{\bot})} \over{{k}_{\| }}} \mbox{\ ln \ } \Biggl| {{2k_F+{k_\|}} \over {2k_F-{k_\|}}} \Biggr| \label{equa4} \end{eqnarray} where ${F({\bf k})}=e^{- k^2 l^2/2}$ and $l$ is the associated magnetic length defined by: ${l^2}=\hbar c/eB_0$. The expansion of $\chi^{-1}$ to second order in $\bf k$ gives: \begin{eqnarray} -{\chi^{-1}(\bf k)}={{4{\pi^4}{\hbar^4}\rho}\over {m^3{\omega_c}^2}}+ {{2{\pi^4}{ \hbar^5}\rho}\over {m^4{\omega_c}^3}}{{\bf k}_{\bot}^2} -{{\hbar^2} \over {12\rho m}}{{k_\|}^2} \label{equa5} \end{eqnarray} so \begin{eqnarray} g_{\bot}={\pi^{4}\hbar^{5}\rho\over m^{4}\omega_{c}^{3}} \end{eqnarray} and \begin{eqnarray} g_{\|}= -{1\over 24 \rho}{\hbar^{2}\over m} \end{eqnarray} The resulting energy functional: \begin{eqnarray} E[\rho] = \int d{\bf r} \, \biggl\{ {{{\hbar\omega_c} \over 2}}\rho+{{{2\pi^4\hbar^4}\over {3m^3{\omega_c}^2}}{\rho}^3}+{{{\pi^4}{\hbar^5}}\over {m^4{\omega_c}^3}}{\rho{\bigl|} {\bf \nabla}_{\bot}\rho {\bigr|}^2}-{{\hbar^2} \over 24m}{{\bigl|}{{\bf \nabla}_{\|}\rho {\bigr|}^2} \over \rho} \biggr\} +\int d{\bf r} \, \rho V_{x} \label{equa6} \end{eqnarray} is the same as the one originally derived by Tomishima and Shinjo [6].\\ Note that even though the free electron states are still plane waves along the field direction the gradient correction for this direction has the opposite sign from the free field case. This is a consequence of the 1-dimensional nature of the high field limit. Explicitly, the free fermion response function in 1-dimension is \begin{eqnarray} \chi(k) ={-m\over \hbar^{2} \pi k} \mbox{\ ln \ } \Biggl| {{2k_F+{k}} \over {2k_F-{k}}} \Biggr| ={-m\over \hbar^{2}\pi k_{F}}\left\{1+{1\over 12}k^{2}/k_{F}^{2}\right\}+O(k^{4}) \end{eqnarray} or \begin{eqnarray} \chi^{-1}(k)={-\hbar^{2}\pi k_{F}\over m}\left\{1-{1\over 12}k^{2}/k_{F}^{2}\right\} +O(k^{4}) \end{eqnarray} with $k_{F}=\pi\rho$. Therefore [13] \begin{eqnarray} g_{\|}=-{\pi\over 24 k_{F}}{\hbar^{2}\over m}=-{1\over 24\rho}{\hbar^{2}\over m} \end{eqnarray} which is exactly the high magnetic field result derived above. \section{Current corrections} The Tomishima and Shinjo approximation is obtained when the variation of vorticity is omitted. However, this variation involves the gradient of $\rho$ and can not a priori be neglected. We therefore generalize our previous expression for $U$ to include vorticity as follows: \begin{eqnarray} U[{\rho},{\bf \nu}] = \int d{\bf r} \, \biggl\{ f(\rho({\bf r}),{\bf \nu({\bf r})})+g_{\bot}(\rho({\bf r}),{\bf \nu({\bf r})}) {\bigl|} {\bf \nabla}_{\bot}\rho {\bigr|}^2+g_{\|}(\rho({\bf r}),{\bf \nu({\bf r})}) {\bigl|} {\bf \nabla}_{\|}\rho {\bigr|}^2 \biggr\} \label{Urho} \end{eqnarray} {}From the definition of the vorticity: \begin{eqnarray} {\bf \nu({\bf r})}=-{{{eB_0}\over mc}{\bf u}_z}+\bf \nabla\times \bigl({\bf j \over \rho}\bigr) \;. \label{vort} \end{eqnarray} Consequently, we do not need to add in the $U[{\rho},{\bf \nu}]$ expansion terms in ${\bigl|} {\bf \nabla\nu} {\bigr|}$ which are higher order. Variation of the energy with respect to ${\bf j}_p$, $\delta E/\delta {\bf j}_p=0$, using the chain rule result \[ {\delta \over \delta {\bf j}_{p}({\bf r})} ={1\over{\rho({\bf r})}} {\bf \nabla}\times {\delta\over{\delta\nu ( {\bf r})}} \] gives: \begin{eqnarray} {\bf j}=-{1 \over m}{\bf \nabla\times} {{\partial f}\over{\partial{\bf \nu}}} \label{cou5} \end{eqnarray} to second order. The Euler equation, $\delta E/\delta\rho({\bf r})=\mu$, becomes: \begin{eqnarray} & &{{\partial f}\over{\partial\rho}}[{\rho},{\nu}] -2g_{\bot} {\bf \nabla}_{\bot}^2\rho -2g_{\|} {\bf \nabla}_{\|}^2\rho \nonumber\\ & & \mbox{\ } -{{\partial g_{\bot}}\over {\partial\rho}} {\bigl|} {\bf \nabla}_{\bot}\rho {\bigr|}^2 -{{\partial g_{\|}}\over {\partial\rho}} {\bigl|} {\bf \nabla}_{\|}\rho {\bigr|}^2=-V_{x}({\bf r})+\mu+{m\over2} {{\bf j\bf j}\over\rho^2} \label{equa15} \end{eqnarray} At uniform density we get:$${{\partial f}\over{\partial\rho}}[{\rho_0},{\nu_0}]=\mu[{\rho_0},{\nu_0}] \mbox{\ \ \ \ with \ \ \ \ } {{\bf \nu}_0}=\omega_{c}\hat{{\bf z}} $$ so the expression for $f$ is formally the same as (\ref{equa2}) with the variables $\rho$ and $\bf \nu$: \begin{eqnarray} f[{\rho},{\bf \nu}]={{\hbar \over 2}|{\bf \nu}|}\rho+{{2\pi^4\hbar^4} \over {3m^3{|{\bf \nu}|}^2}}{\rho}^3 \;. \label{equa9} \end{eqnarray} With equation (\ref{cou5}), we can now calculate the total current density, \begin{eqnarray} {\bf j} {\simeq}- \biggl\{ {{\hbar \over 2m}-{{4\pi^4\hbar^4}\over{m^4}}{\rho^2 \over{|\nu_0|^3}}} \biggr\}{{\bf \nabla}\rho\times {\bf u}_z} \label{equa10} \end{eqnarray} where we have replaced $|\bf \nu|$ by $|\bf \nu_0|$ in the second member of (\ref{equa10}) since the $\bf \nu$ variations are higher order. In the limit of very strong fields we find the result of Skudlarski and Vignale~[14]. On the other hand equation (\ref{equa3}) remains unchanged at $\rho_0$ since all the additional terms are proportional to ${\bf \nabla}\rho$. Thus, the expressions for $g_{\bot}$, $g_{\|}$ are absolutely the same as in section III with ${\omega_c} \to {|\bf \nu|}$. In return, E is modified by the presence of the current, and for a comparison with Tomishima and Shinjo we have to develop $\bf \nu$, eqn. (\ref{vort}), using eqn. (\ref{equa10}): \begin{eqnarray} {|{\bf \nu}|} = {\omega_c} - {\hbar \over 2m}{\biggl[}{{{\bigl|}{{\bf \nabla}_{\bot} \rho{\bigr|}^2} \over \rho^2} - {{{\bf \nabla}_{\bot}^2 \rho} \over \rho}}{\biggr]} - {{{4\pi^4}{\hbar^4}}\over {m^4{\omega_c}^3}}{\biggl[}{{\bigl|}{\bf \nabla}_{\bot} \rho{\bigr|}^2}+{\rho{\bf \nabla}_{\bot}^2 \rho} {\biggr]} \;. \label{equa12} \end{eqnarray} Using this expansion in eqn. (\ref{equa9}) for $f[\rho,\nu]$ and with eqn. (\ref{equa10}) for the current we finally obtain for the energy functional E: \begin{eqnarray} & E[\rho] = \int d{\bf r} \, &\biggl\{ {{{\hbar\omega_c} \over 2}}\rho+{{{2\pi^4\hbar^4}\over {3m^3{\omega_c}^2}}{\rho}^3}-{{\hbar^2}\over 8m}{{\bigl|}{{\bf \nabla}_{\bot}\rho {\bigr|}^2} \over \rho}+{{{3\pi^4}{\hbar^5}}\over {m^4{\omega_c}^3}}{\rho{\bigl|} {\bf \nabla}_{\bot}\rho {\bigr|}^2}\nonumber \\ & &-8m\left({{{\pi^4}{\hbar^4}}\over{m^4{\omega_c}^3}}\right)^{2} \rho^{3}\left|{\bf \nabla}_{\bot}\rho\right|^{2} -{{\hbar^2}\over 24m} {{\bigl|}{{\bf \nabla}_{\|}\rho {\bigr|}^2} \over \rho} \biggr\} +\int d{\bf r} \, \rho V_{x} \label{equa13} \end{eqnarray} \section{Conclusion} Current corrections do not change anything along the field direction since this correction is only due of the 1-dimensional nature of the high field limit. The transverse corrections however, have greatly changed. First we note an important additional term in $|{\bf \nabla}_{\bot}\rho |^{2}/ \rho$ which does not depend on the strong field intensity. Perpendicular gradients are favored by this term. The next term in ${{\bf \nabla}_{\bot}\rho}$, which is small compared to the constant term, is the same as found by Tomishima and Shinjo but with a factor $+3$. This term and the third term in ${{\bf \nabla}_{\bot}\rho}$ decrease with $B_0$. Therefore, contrary to Tomishima and Shinjo the total transverse corrections increase with $B_0$. This result modifies the transverse pinch. In summary, as expected, the behavior of the total corrections produces an anisotropic density profile. This anisotropy is caused by a reduction of dimensionality due to the high field limit. Our approach, which is technically simple, allows us to extend the calculus to finite temperatures and to include some Landau levels. But since the minimization of eqn.~(\ref{equa13}) is rather complicated, we leave its numerical solution for a forthcoming article. \acknowledgements One of us, A.M. would like to thank Dr D. Levesque for his kind hospitality at LPTHE and the CEA for financial support. LPTHE is Unit\'e de Recherche de L'Universit\'e Paris XI associ\'ee au CNRS. Work done at the Lawrence Livermore National Laboratory is supported by the U.S. Department of Energy under Contract W-4705-Eng-48.
\section{Introduction} The Standard Model of electroweak interactions is undeniably one of the most successful theories in particle physics, accommodating all observed phenomena to date. For all its glory, the theory has the undesirable feature that most of its parameters, although physically well motivated, must be determined empirically. As an example, the way in which the weak interaction can mix quarks of different flavor is described by the Cabibbo-Kobayashi-Maskawa matrix\cite{cabibbo,km}. The nine entries in this $3\times 3$ matrix contain information about the relative strengths and phases with which the ($u,c,t$) quarks couple to the ($d,s,b$) quarks, and their values must be measured experimentally. Our present knowledge of the magnitudes of these parameters are summarized below\cite{pdg}. \begin{equation} {\bf\rm V} = \left(\matrix{ V_{ud} & V_{us} & V_{ub} \cr V_{cd} & V_{cs} & V_{cb} \cr V_{td} & V_{ts} & V_{tb} \cr}\right) = \left(\matrix{ 0.9747\rightarrow 0.9759 & 0.218\rightarrow 0.224 & 0.002\rightarrow 0.005 \cr 0.218\rightarrow 0.224 & 0.9738\rightarrow 0.9752 & 0.032\rightarrow 0.048 \cr 0.004\rightarrow 0.015 & 0.030\rightarrow 0.048 & 0.9988\rightarrow 0.9995\cr}\right) \label{ckmb} \end{equation} \noindent Note that the extreme diagonal elements, $V_{ub}$ and $V_{td}$, are the least well determined. \medskip In this paper the latest inclusive and exclusive measurements of the ratio $|V_{ub}/V_{cb}|$~will be discussed. These were both made by the CLEO collaboration, using data accumulated at the Cornell Electron Storage Ring (CESR). In the CLEO-II detector, three concentric tracking devices provide charged-particle momentum resolution of $\sigma_p/p = 0.005+0.0015p$ ($p$ in GeV/c), and a 7800 crystal CsI calorimeter provides neutral shower energy resolution of $\sigma_E/E = 0.019 - 0.001E + 0.0035/E^{0.75}$. More detailed information is available elsewhere\cite{cleodetector}. \smallskip The data sample used in the inclusive analysis consists of $924~pb^{-1}$ accumulated with the CESR center of mass energy tuned to the $\Upsilon$(4S) resonance, and $416~pb^{-1}$ accumulated at energies below the $B\overline{B}$ production threshold. The more recent exclusive analysis was done using approximately a factor of two more data. \section{Inclusive Measurements} All determinations of $V_{ub}$ ~to date have been made by studying charmless semileptonic decays of $B$ mesons produced in $e^+-e^-$ collisions with center of mass energy at or near the $\Upsilon$(4S) resonance \cite{cleo90,argus90,argus91,cleo93,jknelson}. These are all measurements of the inclusive momentum-dependent rate of leptons, $dN_\ell/dP_\ell$, from $B\rightarrow X_u\ell\nu$ decays\cite{leptons}. Since the rate of $b\rightarrow u\ell\nu$ ~is very small, the main experimental challenge in these analyses is the suppression of backgrounds. For inclusive $b\rightarrow u\ell\nu$ ~analyses, the three main sources of unwanted leptons are $b\rightarrow c\ell\nu$~decays, other $B$ meson decays (for example $B\rightarrow\psi X, \psi\rightarrow\ell^+\ell^-$), and continuum (non $B\overline{B}$) events. ``Fake'' leptons, for example $\pi$'s that penetrate the detector iron and are misidentified as $\mu$'s, must also be considered. \subsection{$b\rightarrow c\ell\nu$~Suppression} The elimination of $b\rightarrow c\ell\nu$~decays from the $b\rightarrow u\ell\nu$~sample is achieved with a simple lepton momentum cut. As can be seen in Fig.~\ref{leptonspectrum}(a), the kinematic endpoint momentum of leptons from $B\rightarrow X_c\ell\nu$ is about 2.4 GeV/c\cite{batrest}. Fig.~\ref{leptonspectrum}(b) shows theoretical lepton spectra for several models of $b\rightarrow u\ell\nu$\cite{accmm,isgw,wsb}. Although the models differ significantly in detail, they all share the basic kinematic feature that the lepton momentum endpoint extends to $P_\ell\sim 2.6$ GeV/c. \smallskip The approach taken in all inclusive analyses has been to examine the yield of leptons only in the endpoint region between $\sim 2.4$ GeV/c and $2.6$ GeV/c\cite{lowerpcut}. This restriction eliminates most of the $b\rightarrow c\ell\nu$ contamination, but only at the price of introducing a severe model-dependence in the procedure used to extract $V_{ub}$~from the data. The reason for this is clear upon examination of Fig.~\ref{leptonspectrum}(b). The models shown differ considerably, both in overall rate (plot area) and lepton momentum dependence (plot shape). The effect of these differences are discussed in a later section. \begin{figure}[htb] \vspace{0.in} \begin{center} \unitlength 1.0in \begin{picture}(6.,2.5)(0,0) \put(0.,0.) {\psfig{bbllx=0in,bblly=0in,width=2.8in,file=lepspec.ps}} \put(3.,0.) {\psfig{bbllx=0in,bblly=0in,width=2.8in,file=btouspec.ps}} \end{picture} \medskip \caption{(a) Inclusive lepton momentum spectrum showing the $b\rightarrow c\ell\nu$ ~(dashed line) and $b\to c\to y\ell\nu$ (dot-dashed line) components, as well as the total fit to CLEO data using the ACCMM model (solid line). (b) The predicted $b\rightarrow u\ell\nu$ ~lepton momentum spectra for the ACCMM (solid line), ISGW (dotted line) and WSB (dashed line) models.} \label{leptonspectrum} \end{center} \end{figure} \subsection{Continuum Suppression} Leptons from non-$B\overline{B}$ continuum events, for example $e^+e^-\rightarrow c\overline{c},~ c\rightarrow s\ell\nu$, are not kinematically excluded from the inclusive $b\rightarrow u\ell\nu$ ~signal region between 2.4 and 2.6 GeV/c. This is a large source of background, and is dealt with in two steps. First, a set of continuum suppression requirements are used to eliminate most of these events. Second, the remaining background is removed by subtracting the luminosity scaled lepton spectrum obtained by analyzing ``signal free'' data accumulated at center of mass energies below the $B\overline{B}$ production threshold. \smallskip The most powerful continuum suppression requirements are topological in nature, designed to select the typically spherical $B\overline{B}$ events while rejecting the more ``jetty'' continuum background. The shape variable used by CLEO is the Fox-Wolfram parameter $R_2 = H_2/H_0$\cite{r2}. The missing momentum of an event, $p_{miss}$, provides additional discrimination against continuum processes. This quantity should be large for $b\rightarrow u\ell\nu$ events where the neutrino carries off appreciable momentum. The requirements that $R_2 < 0.2$, that $p_{miss} > 1~$GeV/c, and that the lepton and missing momentum point into opposite hemispheres are used. The net effect of these cuts is to reduce the continuum background by a factor of 70 while maintaining 38\% efficiency for the $b\rightarrow u\ell\nu$ signal. \smallskip The statistical uncertainty introduced by the continuum subtraction is a function of the size of the continuum data sample. The normal operating mode of CESR/CLEO-II is to accumulate data on the $\Upsilon$(4S) resonance two-thirds of the time, and just below $B\overline{B}$ threshold the remaining one-third. The resulting continuum data sample is large enough that the error introduced by the subtraction are much smaller than the statistical error on the signal yield. \subsection{Other Backgrounds} After continuum suppression and subtraction, the events remaining in the lepton momentum endpoint region are either true $b\rightarrow u\ell\nu$ ~signal or non-continuum background. This background is due to leptons from $b\rightarrow c\rightarrow X\ell\nu$, from $B\rightarrow\psi$ and $B\rightarrow\psi^\prime$ decays, and from fake leptons. The fraction of events in the lepton spectrum signal region due to these sources is less than 10\%, is well understood from studying the data, and is carefully accounted for when calculating the final $b\rightarrow u\ell\nu$~yield. \subsection{Signal Extraction and Model Dependence} Fig.~\ref{endpoint}(a) shows lepton spectra from $\Upsilon$(4S) and continuum data. The solid line is a fit to the continuum lepton distribution. Fig.~\ref{endpoint}(b) shows the result of luminosity scaling and subtracting this fitted line from the $\Upsilon$(4S) data, as well as the predicted contribution from $b\rightarrow c\ell\nu$ ~decays. A significant excess of events is seen, indicating the presence of charmless B decays. \begin{figure}[htb] \vspace{0.in} \begin{center} \unitlength 1.0in \begin{picture}(3.,3.)(0,0) \put(-0.5,0.1) {\psfig{bbllx=0pt,bblly=0pt,width=3.0in,height=3.0in,file=inc.ps}} \end{picture} \bigskip \caption{CLEO-II inclusive lepton momentum distributions. Shown in (a) are lepton spectra from $\Upsilon$(4S) (filled circles) and continuum (hatched histogram) data, as well as a fit to the continuum lepton distribution (solid line). The filled circles in (b) show the result of subtracting the fitted and scaled continuum distribution from the $\Upsilon$(4S) data. The solid histogram shows the predicted contribution from $b\rightarrow c\ell\nu$ ~decays.} \label{endpoint} \end{center} \end{figure} Extracting $|V_{ub}/V_{cb}|$~ from the endpoint data involves several model dependent factors: \begin{equation} \left|{V_{ub}\over V_{cb}}\right|={\Delta B_{ub}(p) \over d(p)B_{cb}} \label{extvub} \end{equation} In this expression $B_{cb}$ is the $b\rightarrow c\ell\nu$ ~branching ratio and $\Delta B_{ub}(p)$ is the partial branching ratio observed in the endpoint signal region\cite{pdependent}: \begin{equation} \Delta B_{ub}(p) = {N_{ub}(p)/\epsilon(p) \over 2N_{B\overline{B}}} \label{partial} \end{equation} where $N_{ub}(p)$ is the number of $b\rightarrow u\ell\nu$~events in the signal region, $\epsilon(p)$ is the detection efficiency for these events, and $2N_{B\overline{B}}$ is the number of $B$ mesons produced. $\Delta B_{ub}(p)$ is somewhat model dependent since the detection efficiency $\epsilon$ depends weakly on the shape of the lepton momentum spectrum. \smallskip The term $d(p)$ is the product of several strongly model-dependent factors: \begin{equation} d(p) = f_u(p){\gamma_u \over \gamma_c} \label{scalefact} \end{equation} where $f_u(p)$ is the fraction of the total $b\rightarrow u\ell\nu$~spectrum in the endpoint signal region (model dependence on spectrum shape), and $\gamma_u$ relates the semileptonic width and $V_{ub}$~via $\Gamma_{b\rightarrow u\ell\nu} = \gamma_u |V_{ub}|^2$\cite{gammac}, (model dependence on spectrum area). \smallskip Table \ref{incresults} shows the factors $d(p)$ and $\epsilon(p)$ as well the final extracted CLEO values of $\Delta B_{ub}(p)$ and $|V_{ub}/V_{cb}|$ for various models. The model-dependence of $|V_{ub}/V_{cb}|$ is larger than the experimental uncertainty, and at present is the biggest factor limiting the accuracy of this measurement. \begin{table}[htb] \caption{CLEO-II results for $d(p)$, $\epsilon(p)$, $\Delta B_{ub}(p)$ and $|V_{ub}/V_{cb}|$. The values listed for $d(p)$ and $\epsilon(p)$ were calculated for the momentum range $2.4 < p_\ell({\rm GeV/c}) < 2.6$, and are shown to illustrate the effect of model dependence. The values of $\Delta B_{ub}(p)$ and $|V_{ub}/V_{cb}|$ represent the extended signal region $2.3 < p_\ell({\rm GeV/c}) < 2.6$.} \label{incresults} \begin{tabular}{lcccc} Model & $d(p)$ & $\epsilon(p)$ & $10^6\times\Delta B_{ub}(p)$ & $|V_{ub}/V_{cb}|$ \\ \hline ACCMM\cite{accmm} & $0.12$ & $0.16\pm 0.01$ & $154\pm 22\pm 20$ & $0.076\pm 0.008$ \\ ISGW\cite{isgw} & $0.05$ & $0.21\pm 0.02$ & $121\pm 17\pm 15$ & $0.101\pm 0.010$ \\ WSB\cite{wsb} & $0.11$ & $0.20\pm 0.02$ & $122\pm 17\pm 16$ & $0.073\pm 0.007$ \\ KS\cite{ks} & $0.19$ & $0.22\pm 0.02$ & $115\pm 16\pm 15$ & $0.056\pm 0.006$ \\ \end{tabular} \end{table} \section{Exclusive Measurements} The CLEO collaboration has recently observed the exclusive charmless semileptonic decay mode $B\rightarrow\pi\ell\nu$. Measurement of this and other exclusive channels will provide a new avenue for determining $|V_{ub}/V_{cb}|$ and studying the $b\rightarrow u\ell\nu$ ~form factors, and should be a powerful tool for testing the various available models. \smallskip This section will provide some details of the CLEO $B\rightarrow h\ell\nu$ analysis, where $h$ is $\pi^\pm$, $\pi^0$, $\rho^\pm$, $\rho^0$ or $\omega$. Table \ref{exclpred} lists the theoretical predictions for the partial widths of $B\rightarrow\pi\ell\nu$~ and $B\rightarrow\rho\ell\nu$, and their ratio. Note again that the predicted widths show a strong model-dependence. \begin{table}[htb] \caption{Predictions for the exclusive partial widths $\Gamma(B^0\rightarrow\pi^-\ell\nu)$ and $\Gamma(B^0\rightarrow\rho^-\ell\nu)$.} \label{exclpred} \begin{tabular}{lccc} Model & $\Gamma(B^0\rightarrow\pi^-\ell\nu)$~[$10^{12}|V_{ub}|^2$ sec$^{-1}$] & $\Gamma(B^0\rightarrow\rho^-\ell\nu)$~[$10^{12}|V_{ub}|^2$ sec$^{-1}$] & $\Gamma(B^0\rightarrow\rho^-\ell\nu)/\Gamma(B^0\rightarrow\pi^-\ell\nu)$ \\ \hline ISGW\cite{isgw} & 2.1 & 8.3 & 4.0 \\ WSB\cite{wsb} & 6.3 -- 10.0 & 18.7 -- 42.5 & 3.0 -- 4.3 \\ KS\cite{ks} & 7.25 & 33.0 & 4.6 \\ ISGW II\cite{isgwii} & 9.6 & 14.2 & 1.5 \\ FGM\cite{faustov} & $3.1\pm 0.6$ & $5.7\pm 1.2$ & $1.8\pm 0.5$ \\ \end{tabular} \end{table} \subsection{Neutrino Reconstruction} The difficulty with exclusive reconstruction of semileptonic decays is the lack of knowledge of the undetected neutrino's 4-momentum. Using the large solid-angle coverage of the CLEO-II detector\cite{coverage} to measure the {\it total} momentum and energy of the {\it rest} of the event, CLEO is able to infer $(E_\nu,\vec{p}_\nu)$ of the neutrino from the missing momentum and energy $(E_{miss},\vec{p}_{miss})$ of the event as a whole: \begin{equation} \label{eneutrino} E_\nu\sim E_{miss} = 2E_{beam} - E^{tot}_{meas}~~{\rm and}~~ \vec{p}_\nu\sim \vec{p}_{miss} = - \vec{p}^{~tot}_{meas} \end{equation} where $E^{tot}_{meas}$ ($\vec{p}^{~tot}_{meas}$) is the total measured energy (momentum) of all tracks and showers in the event. Analysis of Monte Carlo generated $B\rightarrow\pi\ell\nu$~events yields resolutions of 110 MeV and 260 MeV for $|\vec{p}_\nu|$ and $E_\nu$ respectively. \smallskip This method assumes the neutrino from $b\rightarrow u\ell\nu$~is the only undetected particle, making it crucial to reject events containing additional unseen particles (neutrinos and/or $K_L$'s). This is accomplished by demanding that candidate events have only one identified lepton, have zero net charge, and that the reconstructed mass of the neutrino ($m_\nu^2=M^2_\nu - P^2_\nu$) be consistent with zero\cite{zeromasscut}. \smallskip Additional constraints are placed on the final state hadrons. Candidate $\pi^0$'s must have a $2\gamma$ invariant mass within $2\sigma$ (about 12 MeV) of the nominal $\pi^0$ mass, and $2\pi$ ($3\pi$) combinations must have invariant mass within 90 MeV (30 MeV) of the $\rho$ ($\omega$) mass. Identified leptons are required to have a momentum greater than 1.5 GeV/c (2.0 GeV/c) in the $\pi\ell\nu$ ($\rho/\omega\ell\nu$) modes. \smallskip For events passing the above neutrino, meson, and lepton selection requirements, the reconstructed ``beam constrained'' $B$ mass $m_B\equiv\sqrt{E^2_{beam}-|\vec{p}_h + \vec{p}_\ell + \vec{p}_\nu|^2}$ and energy difference $\Delta E\equiv E_{beam}-(E_h-E_\ell-|\vec{p}_\nu|)$ are calculated. Real $B\rightarrow h\ell\nu$~events should have $\Delta E$ close to zero and $m_B$ close to the $B$ meson rest mass. Monte Carlo studies are used to determine the optimum location of the signal region in the $\Delta E-m_B$ plane, finding -250 $< \Delta E {\rm (MeV)} <$ 150 and 5.265 $< m_B {\rm (GeV/c^2)} <$ 5.2875. \smallskip As in the inclusive analysis, event shape variables\cite{r2} are used to suppress continuum backgrounds, and a continuum subtraction is done. The contribution from fake leptons is studied using data, and is also subtracted. The remaining background is predominantly due to $b\rightarrow c\ell\nu$~decays containing an additional $\nu$ or $K_L$, and to cross-feed from other $b\rightarrow u\ell\nu$~modes. \begin{figure}[htb] \vspace{0.in} \begin{center} \unitlength 1.0in \begin{picture}(3.,3.)(0,0) \put(-2.0,0.0) {\psfig{bbllx=0pt,bblly=0pt,height=2.5in,file=excl.eps}} \end{picture} \bigskip \caption{Beam constrained mass distributions for combined $\pi\ell\nu$ (left) and $\rho\ell\nu$ (right) modes. The points are continuum and fake subtracted data, the histograms show the contribution from signal (hollow), $b\rightarrow c\ell\nu$~(shaded), and $b\rightarrow u\ell\nu$~cross-feed (hatched).} \label{lawrence} \end{center} \end{figure} \subsection{Signal Extraction and Model Dependence} After subtracting the contributions from continuum and fake leptons and selecting events in the $\Delta E$ signal region, the beam-constrained mass ($m_B$) distributions of the five signal modes are simultaneously fitted. The shapes of the $b\rightarrow c\ell\nu$~and cross-feed background contributions are obtained from Monte Carlo. The normalization of the $b\rightarrow c\ell\nu$~background for each mode are free parameters in the fit. The isospin relations ${1\over 2}\Gamma(B^0\rightarrow\pi^-\ell^+\nu)= \Gamma(B^+\rightarrow\pi^0\ell^+\nu)$ and ${1\over 2}\Gamma(B^0\rightarrow\rho^-\ell^+\nu)= \Gamma(B^+\rightarrow\rho^0\ell^+\nu)\approx \Gamma(B^+\rightarrow\omega\ell^+\nu)$ constrain the relative neutral and charged meson rates. The total signal and cross-feed background for all five modes is in this way parameterized by two numbers $N_\pi$ and $N_\rho$, the total yield of $\pi$ events and $\rho$ events respectively. \smallskip The beam-constrained mass distributions for $B\rightarrow\pi\ell\nu$ ~and $B\rightarrow\rho\ell\nu$ {}~are shown in Fig.~\ref{lawrence}. The results of the fit for the case of $B\rightarrow\pi\ell\nu$~ are summarized in Table \ref{pifit}, where the signal yield, efficiency, $b\rightarrow c\ell\nu$~ background, and $b\rightarrow u\ell\nu$~ cross-feed probabilities are shown for both the ISGW and WSB models. Note that although the yields are very similar for both models, the efficiencies differ significantly. The contribution from higher mass and non-resonant $b\rightarrow u\ell\nu$~modes, denoted ``other $u\ell\nu$'', is fixed by inclusive lepton endpoint spectrum measurements. \smallskip Many checks have been performed to verify that the observed signal is real, including fitting the energy difference ($\Delta E$) distribution rather than the beam-constrained mass, examining the lepton momentum spectrum and the distribution of angles between the pion and the lepton in the $W$ rest-frame. More details of these studies, as well as a discussion of systematic errors, can be found elsewhere\cite{moriond}. \medskip \begin{table}[htb] \caption{Backgrounds, efficiencies and fit results for the $B\rightarrow\pi\ell\nu$~ analysis.} \medskip \label{pifit} \begin{tabular}{lcccc} & \multicolumn{2}{c}{$\pi^-\ell\nu$} & \multicolumn{2}{c}{$\pi^0\ell\nu$} \\ & ISGW & WSB & ISGW & WSB \\ \hline Raw Data & \multicolumn{2}{c}{30} & \multicolumn{2}{c}{15} \\ Continuum Bkg. & \multicolumn{2}{c}{$2.3\pm0.8$} & \multicolumn{2}{c}{$1.0\pm0.5$} \\ Fake Lepton Bkg. & \multicolumn{2}{c}{$1.2\pm0.3$} & \multicolumn{2}{c}{$0.7\pm0.2$} \\ other $u\ell\nu$ Bkg. & \multicolumn{2}{c}{0.6} & \multicolumn{2}{c}{0.2} \\ Efficiency & 2.9\% & 2.1\% & 1.9\% & 1.4\% \\ Signal Yield & $15.6\pm5.3$ & $16.3\pm5.3$ & $5.0\pm1.7$ & $5.3\pm1.7$ \\ $b\to c$ Bkg. & $9.8\pm1.1$ & $9.8\pm1.1$ & $1.8\pm0.5$ & $1.7\pm0.5$ \\ $\rho/\omega$ Bkg.&$3.8\pm1.7$ & $3.4\pm1.4$ & $1.8\pm0.8$ & $1.6\pm0.7$ \\ \end{tabular} \end{table} Correcting the yield for efficiency results in a preliminary measurement of the $B\rightarrow\pi\ell\nu$~ branching ratio: ${\cal B}(B^0\rightarrow\pi^-\ell\nu)= (1.19\pm0.41)\times 10^{-4}$ using ISGW and $(1.70\pm0.55)\times 10^{-4}$ using WSB. The errors shown are statistical only. \smallskip For the vector meson modes the fit results are used to calculate an upper limit rather than a branching ratio since the non-resonant contribution is uncertain. Using the conservative assumption that there is no non-resonant $b\rightarrow u\ell\nu$~background present in the $B\rightarrow\rho\ell\nu$~signal region, the 90\% confidence level upper limits for ${\cal B}(B^0\rightarrow\rho^-\ell\nu)$ is found using the ISGW (WSB) model to be $<3.1\times 10^{-4}$ ($<4.6\times 10^{-4}$), consistent with the upper limits previously published by CLEO\cite{cleoexclusive}. \smallskip One of the more interesting quantities that can be extracted from the data is an upper limit on the {\it ratio} of branching ratios ${\cal B}(B^0\rightarrow\rho^-\ell\nu) / {\cal B}(B^0\rightarrow\pi^-\ell\nu) < 3.4$ at the 90\% confidence level for both WSB and ISGW models. This can be directly compared to the ratio of partial widths found in Table \ref{exclpred}. \section{Conclusions and Future Prospects} At present, the limiting uncertainly in $|V_{ub}/V_{cb}|$ ~is theoretical. The statistical error of the inclusive measurements is about 10\%, and the variation between models is at least twice as large. Even within a single model, variation of parameters within reasonable limits can yield significant changes\cite{hwang}. CLEO is in the process of repeating its inclusive analysis with over than a factor of two more data, which will further decrease the statistical error of the endpoint analysis. \smallskip Experiments can do more, however, to provide guidance to the theoretical community. With sufficient statistics, measuring the $q^2$ distribution of leptons in the inclusive endpoint region should provide useful feedback. The model dependence can in principle be reduced, or at the very least explored, by examining data in different regions of the $q^2-p_\ell$ plane. \smallskip Semileptonic $\Lambda_b$ decays may provide another avenue to study $V_{ub}$. A measurement of the form factors in $\Lambda_c\to \Lambda\ell\nu$ can be related to $\Lambda_b\to p\ell\nu$ using SU(3) and HQET, and used to extract $|V_{ub}/V_{cb}|$\cite{datta}. Several authors have discussed ways of relating the differential spectra for $b\rightarrow u\ell\nu$~ and $b\to s\gamma$ to reduce the uncertainty in the endpoint region due to hadronization\cite{falk,neubert,bigi,korchemsky}. \smallskip The statistical significance of exclusive measurements is still poor, but will improve with experimental running. It is unlikely that exclusive analyses will ever surpass inclusive measurements in terms of raw statistical accuracy, however they will certainly provide a very powerful tool for testing various theoretical models. The experimental limit on the ratio ${\cal B}(B^0\rightarrow\rho^-\ell\nu) / {\cal B}(B^0\rightarrow\pi^-\ell\nu) < 3.4$ at 90\% confidence level is already slightly challenging for some. Other measurements such as lepton momentum spectra, $q^2$ distribution, and vector meson polarization will provide further theoretical tests \cite{faustov2}. Exclusive measurements at low $q^2$ may also prove valuable\cite{akhoury}. It has recently been shown that measurements of $B\rightarrow\rho\ell\nu$ ~and $B\to K^*\ell\overline{\ell}$ can be used to extract $V_{ub}$~ using SU(3) and HQET\cite{sanda}. \smallskip The theoretical problem of determining the form factors needed to calculate $b\rightarrow u\ell\nu$~is also being approached with lattice gauge calculations, and several groups are making progress\cite{martinelli,ukqcd,ape,lanl}. \medskip In conclusion, good theoretical and experimental progress is being made in the quest to determine $|V_{ub}/V_{cb}|$, and the next few years should yield significant advances in both. \bigskip I would like to thank Jeff Nelson, Ron Poling, Lawrence Gibbons and Ed Thorndike for information about the latest CLEO analyses. I would also like to thank John Sloan and Aida El-Khadra for insight regarding lattice calculations, and finally I would like to acknowledge Tom Browder and Klaus Honscheid whose review paper ``B Mesons'' provided valuable information and references. \smallskip \noindent I gratefully acknowledge the support of the Department of Energy and the A. P. Sloan Foundation.
\section{Introduction} \smallskip \par Cosmic strings might be responsible for the formation of large-scale structure which is observed in the Universe today. They are topological defects which form in a phase transition in the early universe. Topological defect models of structure formation are an alternative to inflationary models. In inflationary models one assumes the existence of a scalar field which drives a phase of very rapid expansion in the early universe, and quantum fluctuations produced during this phase turn into classical density perturbations which grow by gravitational instability into present-day structures. In topological defect models the defects, which are present at all times including today (and could possibly be seen directly some day), act as seeds around which structures form. It is important to study theories from both classes of models in order to see which predictions can distinguish between different models, and which predictions are too generic, arising in models based on widely different assumptions. Among the topological defect models, cosmic strings were the first to be investigated in more detail (Brandenberger 1991), recently textures have also been subjected to closer scrutiny (Turok 1991). The cosmic string model looks promising so far, since normalizations of the parameters occuring in the model obtained from observation agree with each other (Bennet, Bouchet \& Stebbins 1992, Perivolaropoulos 1993a). These obervations include galaxy redshift surveys and measurements of the temperature anisotropies in the cosmic microwave background radiation (CMB). Moreover, cosmic strings naturally produce filaments and planar structures in the matter distribution (Silk \& Vilenkin 1984, Vachaspati 1986, Stebbins et al. 1987), in encouraging agreement with recent galaxy redshift surveys. Another nice feature of the cosmic string model is that it works well in the context of hot dark matter (Brandenberger 1987), for which there are candidates known to exist (the $\mu$ and $\tau$ neutrinos), although they may be massless. In contrast, none of the candidates for cold dark matter required in inflationary models has been detected. Topological defect models have not been as closely investigated as inflationary models, and more work is necessary to quantify the predictions of the cosmic string model. It is especially important to find features accessible to observations in which cosmic strings differ from the class of inflationary models, and statistics sensitive to these features. Recently, the probability distribution of peculiar velocities has been put forward as a statistic which might be capable of discriminating between cosmic strings and inflation (Kofman et al. 1994). Here we show that according to the calculations in a simple string toy model, this statistic cannot be used to differentiate between cosmic strings and inflation. The IRAS 1.9 Jy redshift survey was analyzed (Strauss et al. 1990 \& 1992, Yahil 1991) to obtain a uniform galaxy-density map. The peculiar velocities were reconstructed from it via a self-consistent iterative scheme assuming linear biasing between the density fluctuations of galaxies and mass (Nusser et al. 1991). Using these results, Kofman et al. (1994) determined the probability distribution of a single component of peculiar velocities of regions smoothed over several $h^{-1}$ Mpc, with the result that it is consistent with the underlying distribution $p(v_{x})$ being a normal distribution. Their result for the Gaussianity of the velocity distribution is still tentative because of the limited volume sampled, and because velocities were not measured directly but deduced from a redshift survey. Inflationary models predict a normal distribution for $p(v_{x})$, whereas some deviation from gaussianity is expected in the cosmic string model since individual strings impart coherent velocity perturbations over extended regions, as we will see in the following section. The question is, however, just how large this deviation from gaussianity is and whether it is big enough to be detected by current observations. For a certain region receives velocity perturbations from many strings, and by the central limit theorem many nongaussian perturbations can add up to a gaussian signal. Scherrer (1992) has shown that for seed models the velocity field can be very nearly Gaussian even if the density field is nongaussian. In a model where randomly distributed point masses, all of the same mass, accrete matter gravitationally in a universe dominated by hot dark matter, he found that a very low seed density of less than $10^{-2} {\rm{Mpc}}^{-3}$ is required in order for $p(v_{x})$ to be nongaussian. The main aim of this paper is to calculate $p(v_{x})$ within the cosmic string model in order to quantify its departure from gaussianity, and to find out if the IRAS results are also in agreement with this stringy probability distribution $p(v_{x})$. Previously, 3 dimensional rms velocities have been obtained analytically within the cosmic string model (Vachaspati 1992, Perivolaropoulos \& Vachaspati 1994), and numerical simulations have been performed (Vachaspati 1992, Hara, M\"{a}h\"{o}nen \& Miyoshi 1993) to obtain the probability distribution of peculiar velocities, and it was found that the distribution is quite gaussian. Here we present the first analytical calculation of $p(v_{x})$ , within a model for the string network which has been previously employed to make predictions about temperature anisotropies in the CMB (Perivolaropoulos 1993a \& 1993b, Moessner, Perivolaropoulos \& Brandenberger 1994) and the magnitude of peculiar velocities (Vachaspati 1992, Perivolaropoulos \& Vachaspati 1994). The dependence of the deviation from gaussianity on one of the parameters of the model, namely the number $\nu$ of strings per Hubble volume in the strings' scaling solution, is shown explicitly. Our conclusion is that on scales of several $h^{-1}$ Mpc, $p(v_{x})$ from strings deviates only slightly from a normal distribution. On the smallest scales and for the smallest number of strings per Hubble volume, the largest deviation from gaussianity is expected. Performing a $\chi^{2}$ -test of the stringy $p(v_{x})$ for the smallest scale of $6h^{-1}$ Mpc given in Kofman et al. (1994), with imaginary data binned in the same way as in that paper, but drawn from an underlying gaussian distribution (worst case), yields the result that the stringy probability distribution is in agreement with these data. This agreement holds for all values of $\nu$ , the number of long strings per Hubble volume in the strings' scaling solution, including $\nu = 1$. This shows that the hope expressed in Kofman et al. (1994) that the scenario for the formation of large scale structure, where widely separated strings accrete matter in wakes behind them, can be ruled out using the statistic $p(v_{x})$ is not realized. In the next section we will describe the cosmic string model and the mechanism for the production of velocity perturbations. In the third section we will describe the analytic model for the string network and calculate the moment generating function for a single component of the peculiar velocities averaged over several $h^{-1}$ Mpc. Then we will obtain the probability distribution $p(v_{x})$ from this moment generating function, and compare it with observations. \section{Cosmic Strings and Structure Formation} Cosmic strings are linear topological defects formed at a phase transition in the very early universe (Vilenkin 1985). Those originating in the symmetry braking of a Grand Unified Theory possess an enormous mass per unit length $\mu$ ($G \mu \approx 10^{-6}$ , where $G$ is Newton's constant), and they can be responsible for the structures observed today. Strings can have no ends, so they are formed as either infinitely long strings or closed loops. After formation the network of strings quickly evolves towards a scaling solution where the energy density in long strings remains a constant fraction of the total background energy density. This is achieved by intercommutations and self-intersections of the strings leading to the production of small loops, which then decay by emitting gravitational radiation. In this way, some of the energy input into the string network coming from the stretching of the strings due to the expansion of the universe is transferred to the background. Long strings are straight over distances of the order of the horizon, so that the scaling solution can be pictured as having a fixed number $\nu$ of long strings per Hubble volume at any given time. Initially loops were thought to make the dominant contribution to structure formation. At distances much larger than their radius they act as point masses and accrete surrounding matter (Turok \& Brandenberger 1986, Stebbins 1986, Sato 1986). Improved cosmic string evolution simulations (Bennet \& Bouchet 1988, Albrecht \& Turok 1989, Allen \& Shellard 1990) showed that more of the energy density is in long strings, and they are therefore more important, accreting matter in the form of wakes behind them as they move through space (Brandenberger, Perivolaropoulos \& Stebbins 1990, Perivolaropoulos, Brandenberger \& Stebbins 1990). Spacetime around a long straight cosmic string can be pictured as locally flat, but with a deficit angle of $8\pi G \mu$ (Vilenkin 1985). Therefore a string moving relativistically with velocity $v_{s}$ imparts velocity perturbations to surrounding matter towards the plane swept out by the string. If small-scale structure is present on the string, there is in addition a Newtonian force towards the string. The magnitude of this velocity perturbation is given by (Vachaspati \& Vilenkin 1991, Vollick 1992) \begin{equation} u=4 \pi G \mu \gamma_{s} v_{s} f , \;\;\;\;f=1+\frac{1-\rm{T}/\mu}{2(\gamma_{s}v_{s})^{2}} \end{equation} In the absence of small-scale structure on the string,the tension T of the string is equal to its mass per unit length $\mu$, and $f=1$. If small-scale structure is present, $\mu$ denotes the mass per unit length obtained after averaging over the small scale structure, $\rm{T} \ne \mu$, and $f > 1$. We consider the perturbations caused by strings after $t_{eq}$, the time of equal matter and radiation, in a universe filled with hot or cold dark matter. By the present time, the initial velocity perturbation imparted to the dark matter at time $t_{i}$ has grown to (Brandenberger 1987, Stebbins 1987, Hara \& Miyoshi 1990) \begin{equation} u_{i}\approx 0.4u\sqrt{z(t_{i})} \end{equation} Due to compensation (Traschen, Turok \& Brandenberger 1986, Veeraraghavan \& Stebbins 1992), the deficit angle of strings which are straight over a horizon distance, extends out only to a distance of one Hubble radius $H^{-1}$ from the string, so that matter which is farther away does not receive any velocity perturbations. The velocity perturbation given in eq.(1) is independent of distance from the string (up to the Hubble radius), so that cosmic strings impart coherent perturbations over regions of the size of half a Hubble volume. \section{Moment Generating Function} The moment generating function (mgf) $M_{X}(t)$ of a random variable $X$ is defined by \begin{equation} M_{X}(t)=\langle \exp^{tX} \rangle \end{equation} where the brackets denote the ensemble average, and it contains complete information about $X$. In the following, $X = v_{x}$ denotes the random variable for the component of the peculiar velocities smoothed over a region $\cal V$ of comoving radius $R$ in a fixed direction $\hat{e}_{x}$ . We will calculate the mgf of $X$ in order to obtain the moments (and cumulants) and probability distribution $p(v_{x})$ of $X$ from it. The moments of $X$, $\mu_{j} = < X^{j} > $ are given by \begin{equation} \mu_{j}=\left(\frac{d^{j}}{dt^{j}} \right)_{t=0} M_{X}(t) \end{equation} and the cumulants $c_{j}$ are defined by \begin{equation} c_{j}=\left(\frac{d^{j}}{dt^{j}} \right)_{t=0}\ln(M_{X}(t)) \end{equation} The probability distribution $p(v_{x})$ can be expanded in an asymptotic series called Edgeworth series (Scherrer \& Bertschinger 1991, Stuart \& Ord 1987) in terms of the quantities \begin{equation} \lambda_{j} = c_{j}/c_{2}^{j/2} \end{equation} and Hermite polynomials $H_{n}(x)$ defined by \begin{equation} H_{n}(x) = (-1)^{n} e^{x^{2}/2} \frac{d^{n}}{dx^{n}} e^{-x^{2}/2} \end{equation} For distributions with vanishing odd moments as in our case, the expansion is \begin{equation} p(\delta) = \frac{e^{-\delta^{2}/2}}{\sqrt{2} \pi} [1+ \frac{\lambda_{4}}{24} H_{4}(\delta) + \frac{\lambda_{6}}{720}H_{6}(\delta) +\frac{\lambda_{8} + 35 \lambda_{4}^{2}}{40320} H_{8}(\delta) + \cdots ] \end{equation} where $\delta =v_{x}/ \sigma$ , and $\sigma$ is the standard deviation of $X$. Since it is an asymptotic expansion, the remainder is of the order of the last term included (Erdelyi 1956). The mgf has an important property which makes it useful for calculations. For independent random variables $X$ and $Y$, the mgf of the sum is the product of the individual moment generating functions \begin{equation} M_{X+Y}(t) = M_{X}(t) M_{Y}(t) \end{equation} The following calculations are carried out within an analytical model for the string network which has previously been used to obtain the temperature anisotropies in the CMB and the magnitude of peculiar velocities from strings (see references given in the introduction). According to the scaling solution for cosmic strings, there is a fixed number of long strings present per Hubble volume at any given time. After about one expansion time of the universe (Hubble time), $t \rightarrow 2t$, these strings will typically have self-intersected or intercommuted, so that the resulting strings are uncorrelated with the ones at the previous Hubble time. We will assume that during one expansion time $\nu$ long strings move across the Hubble volume. Each string is assumed to be straight over one horizon distance, and the effect of all strings is taken to be the superposition of the effects of the individual strings. The fact that small-scale structure varies along strings is neglected, so that we might somewhat underestimate the degree of non-Gaussianness of the velocity distribution. We also assume that the strings' positions, velocities and orientations at each Hubble time are random and uncorrelated, although this is not strictly true, since - to mention one reason - the string network has the form of a self-avoiding random walk. According to the two previous assumptions, the random variable for the total peculiar velocity, $X$, is the sum of independent random variables for the velocity perturbations from the individual strings. By eq.(9) we can therefore reduce the calculation of the the mgf for $X$ to that of the mgf for the contribution of only one string, and take the products afterwards. In fact we know that there are $\nu$ strings per Hubble volume on average. The products in eq.(9) simplify if we take a Poisson distribution for the number of strings per Hubble volume instead of assuming the presence of exactly $\nu$ of them (Scherrer \& Bertschinger 1993). We can picture this as having a reservoir of $n$ strings, each with a probability $\nu/n$ of being present in a particular Hubble volume, in the limit that $n \rightarrow \infty$. So if $M_{Y_{i}}(t)$ denotes the mgf for a single component of peculiar velocities of the region $\cal V$ due to one string present at time $t_{i}$ in the region's Hubble volume, then \begin{eqnarray} M_{X_{i}}(t) & = & \lim_{n \rightarrow \infty} [ \frac{\nu_{i}}{n} \langle \exp^{tY_{i}} \rangle + ( 1-\frac{\nu_{i}}{n} ) ]^{n} \nonumber \\ & = & \exp{[\nu_{i}(M_{Y_{i}}(t) - 1)]} \end{eqnarray} where $X_{i}$ is the random variable for the contribution of all strings at Hubble time $t_{i}$ to the velocity perturbation of $\cal V$, and $\nu_{i}$ denotes the average number of strings having an effect on $\cal V$ at time $t_{i}$, i.e. those strings which are within a distance of one Hubble radius of $\cal V$ at time $t_{i}$ (see eq.(18)). Since $X = \sum_{i=1}^{N} X_{i}$ , where $N$ is the number of expansion times since $t_{eq}$ \begin{equation} N=\log_{2} \frac{t_{0}}{t_{eq}} \end{equation} and the $X_{i}$ are assumed to be independent, \begin{eqnarray} M_{X}(t) & = & \prod_{i=1}^{N} M_{X_{i}}(t) \nonumber \\ & = & \exp{[ \sum_{i=1}^{N} \nu_{i} ( M_{Y_{i}}(t) - 1)]} \end{eqnarray} We will now calculate the mgf for $Y_{i}$, the random variable for the component in the fixed direction ${\hat{e}}_{x}$ of the peculiar velocities of a region $\cal{V}$ of comoving radius $R$ (the smoothing radius) due to one string affecting the region at time $t_{i}$. We have to perform the ensemble average over positions , orientations ${\hat{e}}_{s}$ and directions of velocity ${\hat{v}}_{s}$ of the string. For a long straight string only transverse velocities are observable, and we assumed ${\hat{e}}_{s}$ and ${\hat{v}}_{s}$ to be random unit vectors. Therefore the unit normal $\hat{e}={\hat{e}}_{s} \times {\hat{v}}_{s}$ of the plane swept out by the string, i.e. the direction in which matter receives velocity perturbations, is itself a random unit vector. Consequently, the projection $s = \hat{e} \cdot {\hat{e}}_{x}$ is uniformly distributed over the interval $[-1,1]$, and the magnitude of the velocity perturbation from one string has to be multiplied by $s$ to get the component in direction ${\hat{e}}_{x}$. During one expansion time a string sweeps out a plane towards which matter within a distance of one Hubble radius receives velocity perturbations, which have grown to $u_{i}$ (see eq.(2)) by today. The possible values $y_{i}$ for the random variable $Y_{i}$, the projection of the peculiar velocity of $\cal{V}$ in direction $\hat{e_{x}}$ due to one string at time $t_{i}$, are a function of the perpendicular distance $r$ of the centre of $\cal{V}$ to this plane: \[ y_{i}(r) = \;s u_{i} r/R_{i} \;\;\;\; {\rm{for}} \;\;\; 0 < r < R_{i}\] \[ y_{i}(r) = \;s u_{i} \;\;\;\; {\rm{for}} \;\;\; R_{i} < r < H^{-1}_{i} - R_{i} \] \begin{equation} y_{i}(r) = \;s u_{i} \frac{H^{-1}_{i}-(r-R_{i})}{2 R_{i}} \;\; {\rm{for}} \;\;H^{-1}_{i}-R_{i}< r<H^{-1}_{i}+R_{i} \end{equation} $R_{i}$ is the physical size of the comoving radius $R$ at time $t_{i}$. Strings within a distance of $H^{-1}(t_{i}) \equiv H^{-1}_{i}$ of the region $\cal V$ can affect it. We distribute the centres $C$ of these planes randomly within a sphere of radius $r_{max}^{i} = H^{-1}_{i}$ around the centre of $\cal{V}$. So the probability $p(c)$ for $C$ to be a distance $c$ from the centre of $\cal{V}$ is \begin{equation} p(c) = \frac{3c^{2}}{(r_{max}^{i})^{3}} \end{equation} Since the normal of this plane has random direction, $r$ can be smaller or equal to $c$, with probability \begin{equation} p(r;c) \approx 2 \frac{r}{c^{2}} \end{equation} Integrating over all $c$ gives the probability for the plane to be a distance r from the centre of $\cal{V}$ as \begin{equation} p(r)= \int_{r}^{r_{max}^{i}} dc \; p(r;c) \; p(c) = \frac{6r}{(r_{max}^{i})^{3}} (r_{max}^{i} -r) \end{equation} The ensemble average thus becomes an integral over $r$ and $s$ \begin{eqnarray} M_{Y_{i}}(t) & = & \langle \exp^{t Y_{i}} \rangle \nonumber \\ & = & \int_{0}^{r_{max}^{i}} dr\;p(r) \int_{-1}^{1} ds \; \frac{1}{2} \exp{(t y_{i}(r))} \end{eqnarray} These integrals can be done, and then eq.(12) can be used to obtain the mgf for $X$ which includes the effect of all strings, with the number of strings affecting $\cal V$ at time $t_{i}$ being given by \begin{equation} \nu_{i} = \nu(r_{max}^{i})^{3}/(H^{-1}_{i})^{3} \end{equation} There is one problem with the above. The formulas for $y_{i}(r)$ in equations (13) are only true if the projection of $\cal{V}$ onto the plane swept out by the string along its normal falls completely into that plane, and is not (partly) outside of it. But the latter can happen for large $c$ for some orientations of the plane, since one side of the plane, in the direction of the string's motion, has a length $l{_i} = H^{-1}_{i}$ or smaller, so that the distance of $C$ to the edge of the plane can be smaller than $H^{-1}_{i} /2$. For $l{_i} = H^{-1}_{i}$ one can show that less than half of the strings miss $\cal{V}$ and give no perturbations to it, so that we can estimate this effect by replacing $\nu$ by $\nu_{\rm{eff}} = 0.5 \nu$ . If the strings are moving slowly, so that $l_{i}$ is even smaller, and not at about the speed of light, our model is not really applicable because the formula for the imparted initial velocity perturbations would change. Actually a string can affect $\cal{V}$ if $c \leq R_{i} + H^{-1}_{i}$ . But for $c$ larger than $H^{-1}_{i}$ we encounter the problem mentioned in the previous paragraph, so that we overestimate the perturbations by using $r_{max}^{i} = H^{-1}_{i} + R_{i}$. Therefore we calculate the cumulants for both $r_{max}^{i} = H^{-1}_{i}$ and $r_{max}^{i} = H^{-1}_{i} + R_{i}$ and take their average, and the model is more accurate for smaller $R$. But since at scales below about $5h^{-1}$ Mpc nonlinear effects become important, and we are only considering linear perturbations, we must also keep above that scale. \section{Probability Distribution and Comparison with Observations} First we want to look at the shape of the probability distribution for $v_{x}$. The nongaussianness is largest on the smallest scales since smoothing makes things more gaussian, and larger regions are affected by more strings. Therefore we are going to compare $p(v_{x})$ from strings with the results from IRAS at the smallest scale of $R = 6 h^{-1}$ Mpc considered in Kofman et al. (1994). Also, the derivation of $M_{X}(t)$ is valid for scales of $R \leq l_{eq}$, where $l_{eq} = 13 h^{-2}$ Mpc is the comoving size of the Hubble radius at the time of equal matter and radiation. The values $\Omega = 1$, $h = 1/2$ and $z_{eq} = 2.3 \cdot 10^{4} \Omega h^{2}$ are used. Using a symbolic manipulation program (O'Dell 1991), the cumulants are obtained from $M_{X}(t)$ according to eq.(5), giving the following values for the $\lambda_{j}$ (eq.(6)) needed in the expansion of the probability distribution (eq.(8)) for $R=6 h^{-1}$ Mpc : \begin{eqnarray} \lambda_{4} & = & 0.34/\nu \nonumber \\ \lambda_{6} & = & 0.20/\nu^{2} \nonumber \\ \lambda_{8} & = & 0.15/\nu^{3} \end{eqnarray} These values, as well as the standard deviations quoted below, are the averages of the two cases $r_{max}^{i}=H^{-1}_{i}$ and $r_{max}^{i}=H^{-1}_{i} + R_{i}$ . For two values of $\nu$, $p(v_{x}/\sigma)$ is plotted in Figure~1, up to the term involving $H_{8}(\delta)$ in the expansion (eq.(8)). For the higher value of $\nu = 10$ strings per Hubble volume, the distribution is practically indistinguishable from a gaussian one. For $\nu = 1$ there is a slight deviation. We want to see if this deviation is significant by performing a $\chi^{2}$ -test with the data given in Kofman et al. (1994), which has been grouped into bins of size $v_{x}/\sigma = 0.25$. The data points fall practically on a gaussian curve, so instead of taking the exact values from the data we calculate the absolute frequencies $m_{j}$ in the $j$ bins expected if the underlying distribution were gaussian. The IRAS 1.9Jy survey maps out a sphere of radius $80 h^{-1} \rm{Mpc}$ , so that there are $(80/6)^{3}$ independent smoothing regions of radius $6h^{-1} \rm{Mpc}$. Let $n_{j}$ be the corresponding absolute frequencies expected from the stringy distribution. Using 20 inner bins, we find $\chi^{2} = 4.63$ for $\nu = 1$, where \begin{equation} \chi^{2} = \sum_{j=1}^{20} \frac{{(m_{j}-n_{j})}^2}{n_{j}} \end{equation} , which is much smaller than the $95 \%$ confidence upper limit of $38.58$ for $19$ degrees of freedom, so that the data is in agreement with the probability distribution from strings. If we replace $\nu$ by $\nu_{\rm{eff}} = \nu / 2$ to take into account that the side of the plane swept out by the string in the direction of its motion is only half the diameter of the Hubble volume, then $\chi^{2} = 22.5$ for $\nu = 1$. Next we want to compare the magnitudes of velocities. The standard deviation of the single velocity components is calculated to be \begin{equation} \sigma = 1.04 \nu^{1/2} \tilde{u} \;\;\; {\rm{for}} \;\;R=6 h^{-1} \rm{Mpc} \end{equation} \begin{equation} \sigma = 0.99 \nu^{1/2} \tilde{u} \;\;\; {\rm{for}} \;\;\;R=12 h^{-1} \rm{Mpc} \end{equation} where $\tilde{u} = 0.4 z_{eq}^{1/2} u$ , and $u$ is defined in eq.(1). We compare these values with results from Peacock and Dodds (1994), who used the power spectra of various observations to calculate the 3 dimensional rms velocities $v_{rms}$ of regions of radius $R$. For a gaussian random variable with three independent gaussian variables as components, the standard deviation of a single component is given by $\sigma = v_{rms}/ \sqrt{3}$. Using this relation, the values given in Peacock and Dodds (1994) are \begin{equation} \sigma = (381 \pm 156) \; {\rm{km/s}} \;\;\; {\rm{for}} \;\;R=6 h^{-1} \rm{Mpc} \end{equation} \begin{equation} \sigma = (337 \pm 138) \; {\rm{km/s}} \;\;\; {\rm{for}} \;\;\;R=12 h^{-1} \rm{Mpc} \end{equation} where we have taken the fractional error of $1/\sqrt{6}$ quoted for the actual measurement of $v_{rms}$ at a scale of $5 h^{-1}$ Mpc. Comparison of $\sigma$ in the string model with the values obtained from observations at these two scales, gives as an average value for $\alpha f$ \begin{equation} \overline{\alpha f} = 3.2 \pm 0.9 \end{equation} where $\alpha$ is the combination of parameters \begin{equation} \alpha= \sqrt{\nu} \frac{G\mu}{10^{-6}} \frac{\gamma_{s} v_{s}}{c} \end{equation} $\alpha$ can be constrained from the rms value of the temperature fluctuations in the cosmic microwave background measured by COBE to be (Perivolaropoulos 1993a) \begin{equation} \alpha = 1.0 \pm 0.2 \end{equation} Using this value of $\alpha$ , we find \begin{equation} \bar{f} = 3.2 \pm 1.1 \end{equation} This indicates that there must be some small scale structure on the strings (see eq.(1)) in order to obtain the right magnitude of peculiar velocities and consistency with CMB observations. This is also in agreement with recent simulations (Bennett \& Bouchet 1988, Albrecht \& Turok 1989, Allen \& Shellard 1990), which show the presence of small-scale structure on cosmic strings. Our analysis has been done in the string wake model, whereas strings with lots of small scale structure accrete matter rather in the form of filaments. Therefore the precise value of $f$ is not to be taken too seriously. Numerical simulations of peculiar velocities from long strings without small scale structure have been performed in a similar framework (Hara, M\"{a}h\"{o}nen \& Miyoshi 1993), where $\nu'$ strings are assumed to move across the horizon at every e-fold expansion of the universe. The authors found that \begin{equation} \frac{G \mu}{10^{-6}} \frac{\gamma_{s}v_{s}}{c} \sqrt{\nu'} =( 4 \pm 1) \end{equation} yields good agreement with observations. The number of strings at every two-fold expansion used in our analysis is related to $\nu'$ by $\nu = \nu'\ln{2}$ . Therefore $\alpha f = 3.3 \pm 0.8 $ from these simulations, which agrees quite well with our result of $\overline{\alpha f} = 3.2 \pm 0.9 $ {}. \section{Discussion} We have shown that the probability distribution of a single component of peculiar velocities in the cosmic string wake model is very close to a normal distribution on scales of several $h^{-1}$ Mpc, as suggested by a general argument for seed models (Scherrer 1992), and in agreement with observations. A comparison of the measured magnitude of peculiar velocities with that expected from strings, using the normalization of string parameters from the COBE quadrupole, suggested that strings have some small-scale structure. Nongaussian features are more apparent in the velocity differences than in the velocities themselves (Catelan \& Scherrer 1994), and it would be interesting to calculate the probability ditribution of these velocity differences within the cosmic string model of structure formation. \section{Acknowledgements} I would like to thank Robert Brandenberger and Leandros Perivolaropoulos for suggestions and helpful discussions. This work was supported in part by the US Department of Energy under Grant DE-FG0291ER40688. \newpage {\bf References} Albrecht A., Turok N., 1989, Phys. Rev. D 40 , 973. Allen B., Shellard E.P.S, 1990, Phys. Rev. Lett. 64, 119. Bennett D., Bouchet F., 1988, Phys. Rev. Lett. 60, 257. Bennett D., Bouchet F., Stebbins A., 1992, ApJ (Lett.) 399, L5. Brandenberger R., 1991, Phys. Scripta T36, 114. Brandenberger R.,Perivolaropoulos L., Stebbins A., 1990, Int. J. Mod. Phys. A5, 1633. Brandenberger R. et al., 1987, Phys. Rev. D36, 335. Catelan P., Scherrer R., 1994, "Velocity Differences as a Probe of Non-Gaussian Density \hspace*{.5in} Fields" (preprint SISSA Ref. 37/94/A). Erdelyi A., 1956, "Asymptotic Expansions", (Dover). Hara T., M\"{a}h\"{o}nen P., Miyoshi S., 1993, ApJ 415, 445. Hara T., Miyoshi S., 1990, Prog. Theor. Phys. 81, 1187. Kofman L. et al., 1994, ApJ 420, 44. Moessner R., Perivolaropoulos L., Brandenberger R., 1994, ApJ 425, 365. Nusser A. et al., 1991, ApJ 379, 6. O'Dell J., 1991, "ALJABR", Fort Pond Research. Peacock J., Dodds S., 1994, MNRAS 267, 1020 . Perivolaropoulos L., 1993a, Phys. Lett. B 298, 305. Perivolaropoulos L., 1993b, Phys. Rev. D 48, 1530. Perivolaropoulos L., Brandenberger R., Stebbins A., 1990, Phys. Rev. D 41, 1764. Perivolaropoulos L., Vachaspati T., 1994, ApJ 423, L77. Sato H., 1986, Prog. Theor. Phys. 75, 1342. Scherrer R., 1992, ApJ 390, 330. Scherrer R., Bertschinger E., 1991, ApJ 381, 349. Silk J., Vilenkin A., 1984, Phys. Rev. Lett. 53, 1700. Stebbins A., 1986, ApJ (Lett.) 303, L21. Stebbins A. et al., 1987, ApJ 322, 1. Strauss M. et al., 1990, ApJ 361, 49. Strauss M. et al., 1992, ApJ 385, 421. Stuart A., Ord J., 1987, "Kendall's Advanced Theory of Statistics", Vol.1 (London: \hspace*{.5in}Charles Griffin). Traschen J., Turok N., Brandenberger R., 1986, Phys. Rev. D 34, 919. Turok N.,1991, Phys. Scripta T36, 135. Turok N., Brandenberger R., 1986, Phys. Rev. D 33, 2175. Vachaspati T., 1986, Phys. Rev. Lett. 57, 1655. Vachaspati T., 1992, Phys. Lett. B, 282, 305. Vachaspati T., Vilenkin A., 1991, Phys. Rev. Lett. 67, 1057. Veeraraghavan S., Stebbins A., 1992, ApJ Lett. 395, L55. Vilenkin A., 1985, Phys. Rep. 121, 263. Vollick D., 1992, Phys. Rev. D 45, 1884. Yahil A. et al., 1991, ApJ 372, 380. \newpage \begin{center} \bf Figure Captions \end{center} Figure 1: Stringy probability distribution $p(\delta= v_{x}/ \sigma)$ of a single velocity component smoothed over regions of radius $R=6 h^{-1}$ Mpc for for $\nu = 1$ (solid line) and $\nu=10$ strings per Hubble volume (dotted line), compared with a normal distribution (dashed line). \end{document}
\section*{References} \begin{list}{}{ \setlength{\parsep}{0pt} \setlength{\itemsep}{0pt} \setlength{\leftmargin}{0pt} } }{ \end{list} } \title{Lognormal Properties of SGR 1806-20 and Implications for Other SGR Sources} \author{K. J. Hurley, B. McBreen, M. Delaney and A. Britton} \date{astro-ph/9508074: presented at 29 ESLAB Symposium, April 1995} \begin{document} \maketitle \begin{abstract} The time interval between successive bursts from SGR 1806-20 and the intensity of these bursts are both consistent with lognormal distributions. Monte Carlo simulations of lognormal burst models with a range of distribution parameters have been investigated. The main conclusions are that while most sources like SGR 1806-20 should be detected in a time interval of 25 years, sources with means about 100 times longer have a probability of about 5\% of being detected in the same interval. A new breed of experiments that operate for long periods are required to search for sources with mean recurrence intervals much longer than SGR 1806-20. \end{abstract} \section{Introduction} The lognormal properties of the soft repeater SGR1806-20 have been previously reported by Hurley, K.J. \etal\ (1994). In particular, both the time interval between repeater events and the luminosity function of the source were fit with lognormal distributions (see Aitchison and Brown, 1957, for a comprehensive introduction to lognormal statistics). This analysis used the data-base of 111 events detected by the International Cometary Explorer (ICE) mission, as reported by Laros \etal\ (1987). While the present number of events observed from the other two sources (Norris \etal, 1991, Kouveliotou \etal, 1993) does not allow any detailed analysis, the intervals between successive events of SGR 0526-66 (Golenetski\v{\i} \etal, 1987) is also suggestive of lognormal behaviour. Continued observations by BATSE of these sources may reveal lognormal properties for one or both of the remaining two repeaters if either passes into a phase of activity similar to the behaviour of SGR1806-20 during 1983. The relationship between the number of active (i.e observable) sources and the true number of SGRs in the galaxy is one which is the subject of some debate (see discussions in Kouveliotou \etal, 1992, Kouveliotou \etal, 1994 and Hurley, K. \etal, 1994). If the time interval between SGR events proves to be lognormal then there may be long quiescent periods where the source could be undetectable, leading to an underestimate of the population. \section{Simulations} In order to investigate the behaviour of sources with much longer mean recurrence times we generated Monte Carlo simulations with a variety of distribution parameters. The Monte Carlo simulations were performed using the random normal generator with Matlab 4.0 for Windows, which is based on a random number generator algorithm given by Park and Miller (1988) with the transformation to the standard normal variate given by Forsythe, Malcolm and Moler (1977). The normal variates were then transformed to lognormal variates using the relationship $Y=e^{\sigma X+\mu}$ where $Y$ is lognormally distributed (with parameters $\mu$\/ and \sig ) and $(\sigma \! X+\mu)$ is normally distributed with mean $\mu$\/ and variance \sig. \begin{figure}[thp] \epsfxsize=\textwidth \epsfbox{mc_1806.eps} \caption{Four separate Monte Carlo simulations of SGR1806-20 activity over a hundred year period, generated from a lognormal distribution of recurrence intervals with the same parameters $\mbox{$\mu_0$}=9.64$ and $\mbox{$\sigma_0$}} \newcommand{\etal}{{\it et al.}=3.44$. The long gaps in activity of the source are the contributions from the tail of this highly skewed distribution.} \label{sgr-sim} \end{figure} The parameters of the lognormal density function which were fit to the distribution of recurrence intervals for SGR1806-20 were $\mbox{$\mu_0$}=9.64$, $\mbox{$\sigma_0$}} \newcommand{\etal}{{\it et al.}=3.44$ (Hurley, K.J. \etal, 1994). Initially we generated 100 year long simulations of SGR1806-20 (Fig.~\ref{sgr-sim}) using these parameters, to check the algorithm. The samples produced were tested for compatibility with a lognormal population using a $\chi^2$ test (Sachs 1986) and were compatible at the 99\% confidence level, indicating that the Monte Carlo simulator was functioning correctly. Two further simulations were then performed to investigate how the source behaviour varied as \mbox{$\mu$}} \newcommand{\sig}{\mbox{$\sigma$} \/ and \sig \/ varied. The results (illustrated in Fig.~\ref{probplot}) are discussed below. \section{Discussion} \begin{figure}[thp] \epsfxsize=\textwidth \epsfbox{probplot.eps} \caption{Percentage of time with one or more events per 25 year interval: (a) as a function of \mbox{$\mu$}} \newcommand{\sig}{\mbox{$\sigma$} ( with $\sigma = 3.0$ fixed) and (b) as a function of \sig (with $\mu=9.5$ fixed). Errorbars are 1 standard deviation values on the mean of 20 runs.} \label{probplot} \end{figure} Presented in Figure~\ref{probplot} are the probabilities for source activity in a 25 year period as calculated from the results of the two simulations described above. Figure~\ref{probplot}(a) shows that as the parameter \mbox{$\mu$}} \newcommand{\sig}{\mbox{$\sigma$} increases the chance of one or more event in 25 years falls from $\approx 80\%$ at $\mu=9.5 \approx \mu_0$ (geometric mean of 0.15 days) to less than 5\% at $\mu=18$ (geometric mean of $\sim \! 500$ days). The chance of one or more events in 25 years for a source like SGR1806-20 (that is with $\mu\approx\mu_0 = 9.64$ and \sig$\approx\sigma_0 = 3.44$) as predicted by Figure~\ref{probplot} indicates that the majority of this type of source should be observed in $\sim$ 25 years. For $\mu\!\gg\!\mu_0$ experiments which operate for a long time must be devised and maintain a continuous search over the whole sky for longer periods than any spaceborne experiments designed so far. Such experiments could reveal a larger population of sources with significant gaps of inactivity. The lognormal distribution arises in statistical processes whose completion depend on a product of probabilities, arising from a combination of independent events (Montroll and Shlesinger, 1982). Lognormal statistics have previously been used in connection with gamma-ray bursts by McBreen \etal\ (1994) and Brock \etal\ (1994). In the context of this investigation the physical significance of this statistical behaviour may lie in the connection between SGRs and neutron stars. In their paper, Hurley, K.J. \etal\ also presented a similar statistical analysis of the behaviour of micro\-glitches from the Vela pulsar (Cordes, Downes and Krause-Polstroff, 1988). The time separation and the intensity of these small ( \(\mid \! \Delta\nu/\nu \! \mid \sim 10^{-9}\) ) frequency adjustments were both compatible with lognormal distributions, and there was no correlation between waiting time and intensity: just as observed with SGR1806-20 (Laros \etal, 1987). This result, combined with the identification of X-ray point sources (Murakami \etal, 1994, Rothschild \etal, 1994) embedded in plerion-powered SNR (Kulkarni \etal, 1993) as counterparts to the SGR sources, suggests structural adjustments in neutron stars may be the cause of SGRs. \section{Conclusion} Previously it was shown that the time intervals between successive events from SGR1806-20 and the associated luminosity function were both lognormally distributed. Structural adjustments in neutron stars may be responsible for this behaviour. The activity of sources with longer mean recurrence times was investigated using Monte Carlo simulations. The results of the simulations indicate that there could exist a significant population of SGRs with means longer than SGR1806-20 that remain undetected. A new breed of experiments with very long observation times will be required to search for this type of source. \begin{refs} \item Aitchison, J. and Brown, J.A.C., 1957, The Lognormal Distribution, Cambridge University Press: Cambridge. \item Brock, M. \etal: 1994, in Fishman, G.J.M, Brainerd, J.J, Hurley, K., ed(s), {\it A.I.P. Conf. Proc. 307,} 672. \item Cordes, J.M., Downs, G.S. and Krause-Polstorff, J., 1988, Astrophys. J., 330, 847. \item Forsythe, G.E., Malcolm, M.A. and Moler C.B., 1977, Computer Methods for Mathematical Computations, Prentice-Hall. \item Golenetski\v{\i}, S.V. \etal, 1987, Sov. Astron. Lett., 13(3), 166. \item Hurley, K. \etal, 1994, Astrophys. J., 423, 709. \item Hurley, K.J., McBreen, B., Rabbette, M. and Steel, S., 1994, Astron. Astrophys., 288, L49. \item Kouveliotou, C. \etal, 1992, Astrophys. J., 392, 179. \item Kouveliotou, C. \etal, 1993, Nature, 362, 728. \item Kouveliotou, C. \etal, 1994, Nature, 368, 125. \item Kulkarni, S.R. \etal, 1994, Nature, 368, 129. \item Laros, J.G. \etal, 1987, Astrophys. J., 320, L111. \item McBreen, B., Hurley, K.J., Long, R. and Metcalfe, L., 1994, MNRAS, 271, 662. \item Montroll, E.W., Shlesinger, M.F., 1982, Proc Nat Acad Sci USA, 79, 3380. \item Murakami, T. \etal, 1994, Nature, 368, 127. \item Norris, J.P., \etal, 1991, Astrophys. J., 366, 240. \item Park, S.K. and Miller, K.W., 1988, Comm ACM, 32(10), 1192. \item Rothschild, R.E., Kulkarni, S.R. and Lingenfelter, R.E., 1994, Nature , 368, 432. \item Sachs, L., 1986, Applied Statistics, Springer-Verlag: New York., \end{refs} \end{document}
\section{introduction} \label{sec:I} The physics of the observed $T_{c}$ suppression in superconductors that contain nonmagnetic disorder has been the subject of much debate in recent years. Let us focus on homogeneously disordered thin superconducting films, with the disorder parametrized by the normal-state sheet resistance $R_{\Box}$, and on the BCS-like quasi transition that is well pronounced in these films although the true superconducting transition is of Kosterlitz-Thouless nature.\cite{films} In these systems, the BCS transition temperature $T_c$, defined as the mid-point of the resistive transition, is observed to decrease monotonically with increasing disorder.\cite{films,R} A complete quantitative understanding of this effect within a microscopic strong-coupling theory has proven difficult, although the first perturbative calculations within a phenomenological BCS model\cite{fukuyama} were rather promising. Qualitatively, disorder-induced changes in the electron-phonon coupling,\cite{schmid} in the Coulomb repulsion between the constituents of the Cooper pairs,\cite{amr} and in the normal-state density of states\cite{dbdos} have all been identified to be important. The difficulty lies in the fact that some of these effects tend to suppress $T_c$ while others tend to enhance it, and $T_c$ depends exponentially on all of them so that subtle balancing effects result. Also, the number of parameters that acquire a disorder dependence is quite large, and fits of theoretical results to experimental data are therefore not necessarily very conclusive. Indeed, theories that are structurally quite different, and mutually inconsistent, have been shown to fit the same $T_c$ data equally well.\cite{f,trkdb} In this situation it is obvious that one should study the disorder dependence of quantities other than the transition temperature in an attempt to discriminate between various theories, and to obtain independent information about the disorder dependence of the various parameters that determine $T_c$. One possibility is to measure the inelastic lifetime of quasiparticles.\cite{qp,pl} The experiment by Pyun and Lemberger\cite{pl} on amorphous InO has been analyzed by the present authors\cite{tpddb} in the framework of a strong-coupling theory for disordered superconductors,\cite{db} and quantitative agreement between theory and experiment has been achieved. Another possibility is to study the influence of nonmagnetic disorder on the pairbreaking induced by a small amount of magnetic impurities in addition to the nonmagnetic ones. This has the advantage that the pair breaking parameter is easier to measure than inelastic lifetimes, and that it can be measured simultaneously with the $T_c$ suppression in a series of films where the nonmagnetic disorder is varied {\it in situ} by controlling the film thickness. Such an experiment has recently been performed by Chervenak and Valles,\cite{jj} who studied quench condensed ultrathin films of Pb$_{0.9}$Bi$_{0.1}$ of varying degrees of disorder ($150\ \Omega < R_{\Box} < 2.2\ {\rm k}\Omega$, leading to a $T_{c}$ between $6$\ K and $2.35$\ K). Of each sample, one half was doped with Gd, while the other half was left undoped. Gd acts as a paramagnetic impurity and leads to pair-breaking and a reduction of $T_{c}$. The transition temperature as a function of Gd concentration was then studied as a function of film thickness, which is correlated with $R_{\Box}$. The remarkable result was that the pair-breaking parameter $\alpha$ is only mildly dependent on disorder for films with normal state sheet resistances $R_{\Box}$ ranging from $0.15\ {\rm k}\Omega$ to $2.2\ {\rm k}\Omega$. The implication seems to be that the effects of disorder that lead to lower transition temperatures do not manifest themselves in the spin-flip pair breaking rate. An attempt to understand this behavior by a phenomenological modification of the Abrikosov-Gorkov result for $\alpha$, using the renormalization of the density of states inherent in Refs.\ \onlinecite{dbdos,db} and \onlinecite{trkdb}, failed.\cite{jj} This poses the important question whether the success of these theories in describing the $T_c$ suppression and the disorder and temperature dependence of the inelastic lifetime was fortuitous, and whether they are lacking some important physical ingredient that manifests itself in the pairbreaking rate. It is the purpose of the present paper to analyze these questions. What we will find is that one needs to take the strong-coupling corrections\cite{am} to the Abrikosov-Gorkov expression into account before one generalizes to the disordered case in order to obtain the correct structure of the theory. Once this is done, we find that our previous theory\cite{db,tpddb} accounts very well for the observed effect. \section{formalism} \label{sec:II} Our starting point is our theory for the suppression of the superconducting $T_c$,\cite{db} and the enhancement of the inelastic scattering rate,\cite{tpddb} by nonmagnetic disorder.\cite{TheoryChoice} In this section we recall the most salient features of this theory. First one uses an exact eigenstate formalism to derive generalized Eliashberg equations for the normal Green function $G(\epsilon,i\omega)$, and the anomalous Green function $F(\epsilon,i\omega)$. Since the wavenumber is not a good quantum number in the presence of static impurities, $G$ and $F$ are functions of energy and frequency, rather than wavevector and frequency as in Eliashberg theory. The Green functions are expressed, as usual, in terms of an anomalous self-energy, $W(\epsilon,i\omega)$, and a normal one.\cite{ssw} The latter is split into a piece $i\omega Z(\epsilon,i\omega)$ that is an odd function of frequency, and a piece $Y(\epsilon,i\omega)$ that is even in $\omega$. In terms of these quantities, $G$ and $F$ read, \begin{mathletters} \label{eqs:1} \begin{equation} G(\epsilon,i\omega)= {i\omega Z(\epsilon,i\omega)+\epsilon+Y(\epsilon,i\omega)\over{[i\omega Z(\epsilon,i\omega)]^{2}-[\epsilon+Y(\epsilon,i\omega)]^{2}}}\quad, \label{eq:1a} \end{equation} and \begin{equation} F(\epsilon,i\omega)={-W(\epsilon,i\omega)\over{[i\omega Z(\epsilon,i\omega)]^2 - [\epsilon+Y(\epsilon,i\omega)]^{2}}}\quad. \label{eq:1b} \end{equation} \end{mathletters}% In clean systems, the normal self-energy piece $Y$ is a constant that just shifts the chemical potential and can be omitted. In the presence of disorder, however, $Y$ has been found to be of crucial importance,\cite{db,blm,db2} and to reflect the physics of the Coulomb anomaly in the density of states\cite{aa} in the context of superconductivity. The generalized Eliashberg equations then take the form of integral equations in both energy and frequency for the three functions $Z$, $W$, and $Y$. A solution of these equations has been obtained by means of various approximations. In particular, the frequency dependence of $Y$ was found to be weak and could be omitted, and its energy dependence was approximated by the first term in a Taylor expansion about a characteristic energy $\bar\omega$, \begin{equation} Y(\epsilon,i\omega) \simeq (\epsilon-\bar\omega) Y^{\prime}\quad, \label{eq:2} \end{equation} with $Y^{\prime} = \partial Y/\partial\epsilon \vert_{\epsilon=\bar\omega}$, and $\bar\omega$ an average phonon frequency.\cite{vdg,db2} With some further approximations, the energy integrations could then be performed, and the theory be cast in the same form as Eliashberg theory. A two-square well approximation then leads to a $T_c$ formula that has the structure of a generalized McMillan or Allen-Dynes formula, viz.\cite{db,db2} \begin{equation} T_c = {\omega_{\log}\over 1.2}\exp\left[{-1.04(1+\tilde\lambda+Y^{\prime}) \over \tilde\lambda - \tilde\mu^{*}[1 + 0.62\tilde\lambda/(1+Y^{\prime})]} \right]\quad. \label{eq:3} \end{equation} Here $Y^{\prime}$ is the normal self-energy piece mentioned above, and $\tilde\lambda$ and $\tilde\mu^{*}$ are disorder dependent generalizations of the electron-phonon coupling constant $\lambda$ and the Coulomb pseudotential $\mu^{*}$, respectively, in Eliashberg theory. Explicit expressions for all three of these quantities have been given in Ref.\ \onlinecite{db}, and will be evaluated for the case of thin films below. In the presence of magnetic impurities, $T_c$ is reduced by pair breaking.\cite{ag} Abrikosov-Gorkov theory has been modified to allow for strong-coupling effects,\cite{am} with the only resulting change being a factor of $1/Z$ in the pair breaking parameter. The result is \begin{equation} -\ln(T_{c}/T_{c0})=\psi(\alpha/2\pi T_{c}+1/2)-\psi(1/2)\quad, \label{eq:4} \end{equation} with T$_{c0}$ the value of $T_c$ in the absence of the magnetic impurities, $\psi$ the di-gamma function, and $\alpha$ the pair breaking parameter $\alpha=(1/Z)(1/\tau_{s})$, with $1/\tau_{s}$ the spin-flip scattering rate. In the case of clean superconductors, $Z=1+\lambda$. In the presence of nonmagnetic disorder, and with the same approximations that lead to the $T_c$ formula given by Eq.\ (\ref{eq:3}), it is straightforward to repeat the calculation of Allen and Mitrovic\cite{am} within the framework of the theory of Ref.\ \onlinecite{db}. The result is again Eq.\ (\ref{eq:4}), but with $\alpha$ replaced by a disorder dependent $\tilde\alpha$ which in turn is related to disorder dependent parameters, \begin{equation} \tilde\alpha={1/\tilde\tau_s\over{1+\tilde\lambda}}\quad. \label{eq:5} \end{equation} Here $\tilde\lambda$ is the same quantity as in Eq.\ (\ref{eq:3}), and $1/\tilde\tau_s$ is the disorder dependent spin flip-scattering rate. Throughout this paper, we choose units such that $\hbar = k_B = 1$. In the next section, we derive an explicit form for $\tilde\alpha$ in a disordered thin superconducting film. \section{disorder dependence of $\alpha$} \label{sec:III} \subsection{Electron-phonon coupling $\tilde\lambda$} \label{subsec:III.A} The electron-phonon coupling strength $\tilde\lambda$ is defined as an integral over the Eliashberg function $\alpha^2 F$, \begin{equation} \tilde\lambda=2\int {d\nu\over{\nu}}\ \alpha^{2}F(\nu)\quad. \label{eq:6} \end{equation} $\alpha^2 F$, and hence $\tilde\lambda$, are disorder dependent due to effects first discussed by Pippard\cite{pippard} in the context of ultrasonic attenuation, and by Schmid\cite{schmid} for the electron-phonon inelastic lifetime. The main physical point is that disorder decreases the coupling of the electrons to longitudinal phonons due to collision drag, but increases the coupling to transverse phonons due to the breakdown of momentum conservation. For realistic parameter values the latter effect is stronger than the former, leading to an overall increase of $\tilde\lambda$ with disorder.\cite{dba2F} For Debye phonons in $3-d$ systems, $\tilde\lambda$ has been calculated in Ref.\ \onlinecite{db}. Repeating that calculation in $d=2$ is straightforward, and very similar to the corresponding calculation of the electron-phonon inelastic lifetime.\cite{dbsds} The result is\cite{tpdthesis} \begin{equation} \tilde\lambda = 2\int_{0}^{\omega_{D}}{d\nu\over{\nu}} {\nu^{2}l\over{\pi m}} \sum_{b=L,T} {d_{b}\over{c_{b}^{3}}}\ f_{b}(\nu l/c_{b})\quad, \label{eq:7} \end{equation} for a system with mean free path $l$. Here $c_{L,T}$ are the longitudinal and transverse speeds of sound, respectively, $\omega_{D}$ is the Debye frequency, the dimensionless constant $d_{b}=k_{F}^{3}/16\pi \rho_{i}c_{b}$ with $\rho_{i}$ the ion density and $k_F$ the Fermi wave number, and the functions $f_{T,L}$ are given by,\cite{dbsds} \begin{mathletters} \label{eqs:8} \begin{equation} f_{T}(x)={8\over{x^{4}}}(1+x^{2}/2-\sqrt{1+x^{2}})\quad, \label{eq:8a} \end{equation} \begin{equation} f_{L}(x)=2\left({1\over{\sqrt{1+x^{2}}-1}} -{2\over{x^{2}}}\right)\quad. \label{eq:8b} \end{equation} \end{mathletters} Substituting Eqs.\ (\ref{eqs:8}) into Eq.\ (\ref{eq:7}) we obtain the disorder dependence of $\tilde\lambda$, \begin{mathletters} \label{eqs:9} \begin{eqnarray} \tilde\lambda&=&{\lambda\hat R_{\Box}E_{F} c_{L}\over{\pi\omega_{D}v_{F}}} \biggl[F_{L}\left({\pi\omega_{D}v_{F}\over{\hat R_{\Box}E_{F}c_{L}}}\right) \nonumber\\ &+&2{c_{L}^{2}\over{c_{T}^{2}}} F_{T}\left({\pi\omega_{D}v_{F}\over {\hat R_{\Box}E_{F}c_{T}}}\right)\biggr] \quad, \label{eq:9a} \end{eqnarray} where we have defined two functions, \begin{eqnarray} &F_{L}(x)&=\sqrt{1+x^{2}}-1-\ln[(\sqrt{1+x^{2}}+1)/2]\quad, \nonumber \\ &F_{T}(x)&= {1-\sqrt{1+x^{2}}\over{2(1+\sqrt{1+x^{2}})}} +\ln[(\sqrt{1+x^{2}}+1)/2]\quad, \label{eq:9b} \end{eqnarray} \end{mathletters}% with $\lambda = 4\omega_{D}d_{L}/\pi m c_{L}^{2}$ the electron-phonon coupling in a clean $2-d$ system. The dimensionless resistance $\hat R_{\Box} = R_{\Box} e^2/\hbar \approx R_{\Box}/4.1\ {\rm k}\Omega$ is a measure of the disorder in the material. As in three dimensions, the size of the disorder renormalization of $\lambda$ depends on the ratio of the longitudinal to the transverse speed of sound. This is a result of the abovementioned competition between an increase in the coupling between electrons and transverse phonons and a decrease of the coupling to longitudinal phonons. Since the transverse speed of sound is invariably smaller than the longitudinal one, $\tilde\lambda$ increases with increasing disorder. This effect tends to reduce the pair breaking rate, Eq. (5). However, we also have to calculate the disorder dependence of the spin-flip scattering rate in order to obtain the disorder dependence of $\tilde\alpha$. \subsection{Spin-flip scattering rate $1/\tau_{s}$} \label{subsec:III.B} The interaction between the electron spin and an impurity spin ${\vec S}({\vec r})$ at site ${\vec r}$ is described by a Hamiltonian, \begin{equation} H_{S}=\sum_{{\bf k},{\bf k^{\prime}},\mu,\nu} J_{{\bf k},{\bf k^{\prime}}}\ {\vec S}({\bf k}-{\bf k^{\prime}})\cdot (c^{\dagger}_{{\bf k^{\prime}}\mu}\vec\sigma_{\mu\nu}c_{{\bf k}\nu})\quad. \label{eq:10} \end{equation} Here $c^{\dagger}$ and $c$ are fermion operators, $\vec\sigma = (\sigma_x, \sigma_y, \sigma_z)$ denotes the Pauli matrices, and $J_{{\bf k},{\bf k^{\prime}}}$ denotes the electron-magnetic impurity exchange integral. We now calculate the electron self energy contribution, $\Sigma$, due to this interaction in Born approximation. It is most convenient to do this in an exact eigenstate representation, in analogy to the calculation of the Coulomb self energy in Ref.\ \onlinecite{aalr}. The calculation is straightforward, and we obtain \begin{equation} \Sigma(\epsilon,i\omega)=\int {d\epsilon^{\prime}\over{N_{F}}} G(\epsilon^{\prime},i\omega)\sum_{{\bf q}}V_{S}({\bf q}) F_{s}({\bf q},\epsilon-\epsilon^{\prime})\quad. \label{eq:11} \end{equation} Here, $N_{F}$ is the free electron density of states per spin at the Fermi level. We only retain the $s-$wave component of the interaction so that $V_{S}({\bf q})=n_{P}S(S+1) J^2$, where $n_{P}$ is the concentration of paramagetic impurities, and J is a measure of the exchange interaction strength.\cite{rkky} $G(\epsilon,i\omega)$ is the normal Green function in the superconductor and is given by Eq.\ (\ref{eq:1a}). Finally, $F_s$ is the spin density analogue of the density-density correlation function denoted by $F$ in Ref.\ \onlinecite{aalr}. If we work to lowest order in the electron-impurity spin interaction, and neglect Coulomb and finite temperature effects in $\Sigma$, then $F$ and $F_s$ are identical. We obtain the spin-flip scattering rate $1/\tau_s$ from the self energy $\Sigma$ by analytically continuing to real frequencies, $i\omega \rightarrow \omega + i0$, and going `on-shell', i.e. putting $\epsilon = \omega$. For our purposes, we are interested in the influence of spin-flip scattering on the superconducting $T_c$. The physics that determines the latter is dominated by processes on a frequency scale of $\bar\omega$, a typical phonon frequency. For the same reason for which we take the parameter $Y^{\prime}$ in Eq.\ (\ref{eq:2}) at the frequency $\bar\omega$ we therefore define \begin{equation} 1/\tau_s = -2{\rm Im}\Sigma(\bar\omega,i\omega\rightarrow\bar\omega + i0) \quad. \label{eq:12} \end{equation} In a clean system, the spin-density correlation function $F_s(q,\omega)$ is frequency independent. In that case we recover from Eq.\ (\ref{eq:12}) the well known result\cite{ag,am} \begin{equation} {1\over\tau_s}=n_{P}S(S+1) J^2 4N_{F}\quad. \label{eq:13} \end{equation} In a disordered system, $F_s$ is diffusive,\cite{aalr} and in the Green function $G$ we must keep the self energy piece $Y^{\prime}$ as discussed above. We thus obtain \begin{mathletters} \label{eqs:14} \begin{equation} {1\over\tilde\tau_s}={2 n_{P}S(S+1) J^2\over{N_{F}[1+Y^{\prime}]}}\sum_{\bf q} F_s\left({\bf q},\bar\omega{\tilde\lambda \over 1 + Y^{\prime}} \right)\quad, \label{eq:14a} \end{equation} where \begin{equation} F_s({\bf q},\omega) = g(q) {Dq^2\over \omega^2 + (Dq^2)^2}\quad, \label{eq:14b} \end{equation} \end{mathletters}% with $g(q)$ the Lindhard function, which for simplicity we replace by $N_F \Theta(2k_F - q)$. Here $D$ denotes the normal phase spin density diffusion coefficient, which in the noninteracting electron approximation coincides with the mass or charge diffusion coefficient, so $D=\pi/m\hat R_{\Box}$. Performing the wavenumber integral in Eq.\ (\ref{eq:14a}) we finally obtain \begin{mathletters} \label{eqs:15} \begin{equation} {1\over\tilde\tau_s} = {1\over\tau_s} \left\{1+{1\over1+Y^{\prime}} {\hat R_{\Box}\over{8\pi}} \ln\left[1 + \left({8\pi\over\hat R_{\Box}}{\epsilon_F\over \bar\omega^{*}}\right)^{2}\right]\right\}\quad, \label{eq:15a} \end{equation} with \begin{equation} \bar\omega^{*} = \bar\omega{\tilde\lambda\over 1 + Y^{\prime}} \quad, \label{eq:15b} \end{equation} \end{mathletters}% and $1/\tau_s$ given by Eq.\ (\ref{eq:13}). $1/\tilde\tau_s$ depends on disorder both explicitly, and implicitly through $Y^{\prime}$. Our final task is therefore to calculate the dependence of $Y^{\prime}$ on $\hat R_{\Box}$. \subsection{Normal Self Energy Piece $Y^{\prime}$} \label{subsec:III.C} In order to calculate $Y^{\prime}$ we again have to repeat the calculations of Ref.\ \onlinecite{db} in $d=2$. Both the electron-electron and the electron-phonon contributions to the self energy contribute to the self energy piece $Y$. Performing a Taylor series expansion in energy around $\epsilon=\bar\omega$ of Eq. (2.12) of Ref.\ \onlinecite{db}, we obtain \begin{equation} Y^{\prime}(\bar\omega)=\delta U_{C}^{Y}(\bar\omega)+4\int {d\nu\over{\nu}} \delta\alpha^{2}F^{H}(\bar\omega,\nu)\quad. \label{eq:16} \end{equation} $\delta U_{C}^{Y}(\bar\omega)$, which describes the Coulomb contribution to $Y^{\prime}$, is taken from Ref.\ \onlinecite{db}, \begin{eqnarray} & &\delta U_{C}^{Y}(\bar\omega)={1\over{\pi N_{F}}}\sum_{\bf q} g({\bf q}){Dq^{2}\over{(Dq^{2})^{2}+\bar\omega^{2}}}\times \label{eq:17} \\ & &\left\{ V_{C}({\bf q})-{2\over{g({\bf q})^{2}}}\sum_{\bf k,p} \sum_{\bf k^{\prime},p^{\prime}}g_{\bf k,k^{\prime}}({\bf q}) g_{\bf p,p^{\prime}}({\bf q})V_{C}({\bf k-p})\right\},\nonumber \end{eqnarray} with the statically screened Coulomb potential, \begin{equation} V_{C}({\bf q})={1\over{2N_{F}}}{\kappa\over{\kappa+q}}\quad, \ \ \ \ \ \ \hfil \kappa=4\pi e^{2}N_{F}\quad. \label{eq:18} \end{equation} Using the prescription for performing momentum sums of this type as described in Ref.\ \onlinecite{db}, the integrals can be done and yield \begin{eqnarray} \delta U_{C}^{Y}(\bar\omega)&=&{\mu \hat R_{\Box}\over{8\pi}} \biggl[G\left({\hat R_{\Box}\bar\omega\over{8\pi E_{F}}}{4k_{F}^{2}\over{\kappa^{2}}},{\hat R_{\Box}\bar\omega\over{8\pi E_{F}}},{2k_{F}\over{\kappa}}\right)\nonumber \\&-&{2\over{\pi}} H\left({\hat R_{\Box}\bar\omega\over{8\pi E_{F}}},{2k_{F}\over{\kappa}}\right)\biggr]\quad, \label{eq:19} \end{eqnarray} with the functions \begin{eqnarray} &G&(x,y,z)={z\over{1+x^{2}}}{1\over{\ln(1+z)}} \nonumber\\ &\times&\biggl\{\ln\left[ {1+1/y^{2}\over{(1+z)^{4}}}\right]-\sqrt{{x\over{2}}}(1-x)\ln\left[ {1-\sqrt{2y}+y\over{1+\sqrt{2y}+y}}\right]\nonumber \\ &+&\sqrt{2x}(1+x)\tan^{-1}\left({\sqrt{2y}\over{y-1}}\right) -2x\tan^{-1}(1/y)\biggr\}\quad; \nonumber \\ &H&(y,z)={z\over{\ln(1+z)}}\ln(1+1/y^{2}) {1\over{\sqrt{z^{2}-1}}} \nonumber\\ &\times &\ln\left[{z+\sqrt{z^{2}-1}\over{z-\sqrt{z^{2}-1}}}\right]\quad. \label{eq:20} \end{eqnarray} The Coulomb pseudopotential $\mu$ in $d=2$ is given by \begin{equation} \mu={\kappa\over{2\pi k_{F}}}\ln\left(1+{2k_{F}\over{\kappa}}\right)\quad. \label{eq:21} \end{equation} As discussed in Ref.\ \onlinecite{db}, the phonon contribution to $Y^{\prime}$, which is given by the second term on the right-hand side of Eq.\ (\ref{eq:16}), is related to a stress-stress correlation function and can be calculated in a similar manner as $\delta U_{C}^{Y}$. One obtains \begin{eqnarray} &4&\int {d\nu\over{\nu}} \delta\alpha^{2}F^{H}(\bar\omega,\nu)=\\ & &\lambda\hat R_{\Box}{c_{L}E_{F}\over{2\pi^{2}v_{F}\omega_{D}}} \sin^{-1}\left[{\omega_{D}v_{F}\over{4 E_{F}c_{L}}}\right] \ln\left[1+\left({8\pi E_{F}\over{\hat R_{\Box}\bar\omega}}\right)^{2}\right]\quad. \nonumber \end{eqnarray} Finally, both contributions can be collected to give \begin{eqnarray} & &Y^{\prime}(\bar\omega)=\hat R_{\Box}\biggl\{{\mu\over{8\pi}} [G-{2\over{\pi}} H] \nonumber \\ &+&\lambda{c_{L}E_{F}\over{2\pi^{2}v_{F}\omega_{D}}} \sin^{-1}\left[{\omega_{D}v_{F}\over{4E_{F}c_{L}}}\right] \ln\left[1+\left({8\pi E_{F}\over{\hat R_{\Box}\bar\omega}}\right)^{2}\right]\biggr\}\ , \label{eq:23} \end{eqnarray} where $G$ and $H$ have the same arguments as in Eq.\ (\ref{eq:19}). The enhancement of $Y^{\prime}$ with increasing disorder is due to the opening of a correlation gap in the (normal state) density of states, and contributes to the decrease of T$_{c}$.\cite{db,db2} \section{Final Result and Discussion} \label{sec:IV} We are now in position to collect our results and thus obtain the disorder dependence of the magnetic pair breaking rate. Substituting Eqs.\ (\ref{eqs:9}), (\ref{eqs:15}), and (\ref{eq:23}) into Eq.\ (\ref{eq:5}) yields our final result, \begin{equation} {\tilde\alpha\over\alpha}= {1+\lambda\over{1+\tilde\lambda}} \left\{1+{1\over{1+Y^{\prime} }}{\hat R_{\Box}\over{8\pi}}\ln\left[1+\left({8\pi E_{F}\over{\hat R_{\Box}\bar\omega^{*}}}\right)^{2}\right]\right\}. \end{equation} The disorder renormalizations of the pair breaking rate appear both in the numerator and denominator and therefore the rate may either increase or decrease with increasing disorder depending upon the material parameters $\lambda,c_{L},c_{T},v_{F},E_{F},\omega_{D},$ and $\mu$. We now address the experiment on Pb$_{0.9}$Bi$_{0.1}$ by Chervenak and Valles.\cite{jj}. To estimate the parameters entering into Eq. (24), let us first consider the parameter values for bulk PbBi, as far as available, given in Ref.\ \onlinecite{ad}. Thereby we have $\bar\omega=56 K$, and $\omega_{D}=108K$. The Bohm-Staver relation gives \begin{equation} {\omega_{D}\over{E_{F}}} {v_{F}\over{c_{L}}}=2 (2/Z)^{1/3}\quad. \label{eq:25} \end{equation} For clean bulk Pb one has $E_{F}=1.1 \times 10^{5} K$, $c_{L}=$2050 m/s, $ c_{T}=710$ m/s, and $Z=4$. However, we do not expect the actual parameter values to correspond to those for either bulk Pb or bulk Pb$_{0.9}$Bi$_{0.1}$. First of all, the material in question is a thin film, and moreover the substrate is expected to modify its properties, in particular the acoustic ones. Evidence for this is provided by the fact that the parameter values quoted above do not give the correct value of the clean limit $T_{c}$ as measured in Ref.\ \onlinecite{jj}. It is therefore likely that the substrate on which the thin layer of Pb is deposited strongly affects the phonon spectra of the film, altering both $\lambda$ and the ratio $c_{L}/c_{T}$ compared to bulk Pb. Accordingly, we choose a value for the bare $\lambda=1.12$ and $\mu=0.1$ which (in the absence of disorder) reproduces the highest T$_{c}$ as measured in Ref.\ \onlinecite{jj}. Lastly, since $c_{L}/c_{T}$ is not known even for Pb$_{0.9}$Bi$_{0.1}$, we let $c_{L}/c_{L}$ be determined by a fit to the data. These parameters provide the curves shown in Figure 1. The solid curves are the results for the disorder dependence of the normalized pair breaking rate as given by Eq. (5). The points represent the data taken from Ref. \ \onlinecite{jj} for two different runs. The decrease at small $\hat R_{\Box}$ is due to the fact that for small disorder, $\tilde\lambda$ and $Y^{\prime}$ grow more rapidly than $1/\tilde\tau_{s}$. The normalized rate goes through a shallow minimum at roughly $\hat R_{\Box}\sim 0.3$ at which point the disorder renormalizations of $Y^{\prime},\tilde\lambda,$ and $1/\tilde\tau_{s}$ are balanced and offset each other. With further increasing disorder the enhancement of $1/\tilde\tau_s$ dominates, and leads to a slowly increasing pair breaking rate. We remark that the point at which the minimum occurs depends sensitively on the ratio of the longitudinal and transverse speeds of sound. To obtain the solid lines in Fig. 1a, $c_{L}/c_{T}=1.9$ was used while 2.1 was used for Fig. 1b. These values lie between the value for bulk Pb (2.88) and the substrate (similar to pyrex, 1.72) used in Ref. \ \onlinecite{jj} and thus does not seem unreasonable. Larger values of $c_{L}/c_{T}$ yield a more drastic reduction of the rate for small disorder and the region of increasing $\alpha$ occurs at larger values of $R_{\Box}$. This sensitivity of the overall shape of the curve to the material parameters may be reflected in the slightly different results obtained for the two experimental runs in Ref. \ \onlinecite{jj} as shown in Fig. 1. It would therefore be very interesting to repeat the experiments using substrates with different acoustic properties. In summary, we have presented a theory for the paramagnetic pair breaking rate in disordered superconducting films and have shown that the disorder dependence of the rate depends delicately on the disorder renormalizations of $Y^{\prime}$, $\tilde\lambda,$ and $1/\tilde\tau_{s}$. As a result, the rate can either increase or decrease with disorder, depending upon material parameters, and in general it is not a monotonic function of disorder. Our conclusion is that the disorder dependence of the rate as observed by Chervenak and Valles in Pb$_{0.9}$Bi$_{0.1}$ films can be quantitatively understood via an application of the microscopic theories developed in Refs. \ \onlinecite{tpddb} and \ \onlinecite{db}. \acknowledgements We gratefully acknowledge helpful discussions with Jay Chervenak, Jim Valles, and Martin Wybourne. Part of this work was performed at the TSRC in Telluride, CO, and we thank the Center for its hospitality. This work was supported by the NSF under grant numbers DMR-92-06023, DMR-92-09879, and DMR-95-10185.
\section{Introduction} \indent Quantum chromodynamics (QCD) tells us that most, if not all, of light hadron masses are generated spontaneously by the breaking of chiral symmetry from {$SU(N_f)\times SU(N_f)$} to diagonal $SU(N_f)$ where $N_f$ is the number of flavors, equal to 2 without strangeness and 3 with. It is also widely believed that as a hadronic system is heated to high temperature or compressed to high density, the broken symmetry will get restored in a way paralleling what happens in condensed matter physics. A natural consequence of the restoration of the chiral symmetry must then be that the spontaneously generated masses disappear as density (and perhaps also temperature) is increased. The question we are raising is how does this ``shedding of mass" occur? This question is at the core of the fundamental theory of matter: How is the mass generated, starting with the lightest object like neutrinos to the heaviest detected particle like the top quark? The aim of this talk is to describe how hadron properties get modified in medium as the system is heated or compressed. That is, immerse a hadron in medium and compress the system or heat it. What does one expect to see happening? To answer this question, let me start with the simplest nuclear system, namely the deuterium. Let us look at what happens when a soft photon is sent in to probe the system. Consider therefore the well-known inverse process \begin{eqnarray} n + p\rightarrow d +\gamma\label{npcap} \end{eqnarray} at thermal neutron energy. This process was first explained in a quantitative way by Riska and Brown \cite{riskabrown} in 1973. What I will do here is to describe it more accurately in a modern QCD framework. Since it is a very low-energy process, QCD can be represented by an effective chiral Lagrangian field theory. This is because at long wavelength limit, chiral perturbation theory (ChPT) is believed to be exactly equivalent to QCD \cite{chptqcd}. This invites us to attempt to describe (\ref{npcap}) in terms of a chiral Lagrangian. There is one basic problem in doing this and that has to do with the description of the deuteron in QCD: we do not really know how to derive the deuteron starting {\it directly} from QCD. Since we are focusing on the chiral aspect of the problem, however, the solution might be sought in a chiral Lagrangian approach to the structure of the deuteron. If the number of colors $N_c$ is in some sense big -- which gives rise to what is known as the ``large $N_c$ limit" -- then the effective Lagrangian is given by meson fields only as we know from the skyrmion structure. There is some important progress in obtaining a bound deuteron, recently through the work of Manton \cite{manton}, as a baryon number 2 skyrmion, but we are still far from understanding it quantitatively. However there is an indirect approach to this which is consistent with QCD and which has the potential to be quantitatively accurate, namely that as we have known all along, the deuteron is made up of a proton and a neutron bound by meson exchanges: In the framework of QCD, the nucleon may be gotten from a large $N_c$ Lagrangian as a soliton (skyrmion) but this is now known to be equivalent, at least in the large $N_c$ limit, to having the nucleon as a matter field in the chiral Lagrangian. We are thus led to consider a chiral Lagrangian that contains baryons (nucleon, $\Delta$ etc.), pseudo-Goldstone bosons $\pi^i$ (pions, kaons etc.), vectors $V_\mu= \omega_\mu, \rho_\mu, \cdots$ etc. with suitable chiral invariant couplings. There have been some attempts to compute the deuteron from such a Lagrangian in chiral perturbation theory \cite{vankolck} but at present the calculation can be done only at low orders since higher order calculations would involve too many parameters to be completely determined from available experiments. Luckily for our purpose, we need not compute the deuteron from first principles as I shall argue below. The important point to note is that certain aspects of the deuteron which have to do with the chiral symmetry structure of hadrons can be probed by the process (\ref{npcap}) {\it without knowing} how to get the nucleus itself from a chiral perturbation theory, as recently discussed by Park, Min and Rho \cite{pmr}. Briefly the argument goes as follows: For physics with energy scale much less than the chiral scale $\Lambda_\chi \sim 4\pi f_\pi\sim 1$ GeV (where $f_\pi$ is the pion decay constant $\sim 93$ MeV), the relevant Lagrangian is, schematically, \begin{eqnarray} {\cal L}=\sum_i {\cal L}_i [B, U, {\cal M}] \end{eqnarray} where $B_i$ are the baryon fields (both octet and decuplet), $U$ the unitary pseudo-Goldstone fields $U=e^{i\pi/f_\pi}$ and ${\cal M}$ the quark mass matrix. The degrees of freedom more massive than the chiral scale $\Lambda_\chi$ are integrated out, appearing implicitly in the counter terms of the Lagrangian. Chiral symmetry requires that there be only derivative couplings apart from terms involving the mass matrix and hence effectively the Lagrangian is an expansion in $\partial/\Lambda_\chi$ and ${\cal M}/\Lambda_\chi^2$. Since the baryon mass is $\sim \Lambda_\chi$, the baryons should be introduced as static matter fields, so the derivative on the baryon field does not involve time derivatives. ChPT is then a systematic perturbative expansion in powers of $Q$, say, $Q^n$, where $Q$ is the energy-momentum scale being probed, with suitable counter terms to remove divergences and to take account of the degrees of freedom that are integrated out. Now in applying this theory to nuclear systems, we need to separate the class of Feynman diagrams into two, one ``irreducible" and the other ``reducible." It is in calculating the irreducible diagrams that ChPT enters. The reducible diagrams -- that cannot be treated by ChPT because of infrared singularity -- are incorporated by solving a Lippman-Schwinger equation or Schr\"odinger equation with the potential obtained with the irreducible graphs by ChPT. This is how bound states are to be treated in ChPT. Now in calculating the process (\ref{npcap}), we can write the EM current in two terms, one the single-particle current $j^{(1)}$ and the other the two-body current $j^{(2)}$ \begin{eqnarray} j_\mu^{EM}=j_\mu^{(1)} + j_\mu^{(2)}.\label{current} \end{eqnarray} The former is called ``impulse approximation current" and the latter ``exchange current." For the system considered, we terminate with the two-body current. Later we will see that in heavy nuclei there can enter many-body currents, some of which become quite important. Very accurate wave functions for the final deuteron and the initial neutron-proton system obtained from some accurate phenomenological potential such as the Argonne $v_{18}$ potential \cite{v18} (or a potential calculated in a high order ChPT if it is feasible) would correspond to a high order chiral expansion since the Schr\"odinger equation sums a certain class of chiral series to all orders and presumably the phenomenological potential also subsumes all orders of chiral perturbation. Now the idea is to compute the matrix element of the current (\ref{current}) in chiral perturbation expansion in such a way that is consistent with the calculation of the wave functions. The calculation of the one-body current $j^{(1)}$ is without ambiguity. That of the two-body exchange current is somewhat subtle, requiring a careful sorting of irreducible and reducible contributions such that the reducible ones are suitably accounted for in the one-body term with the accurate wave functions. In this way, the Ward identities associated with the conserved vector current are satisfied to a given chiral order in the EM current. This is a numerically accurate procedure. \begin{figure} \centerline{\epsfig{file=pmr3fig6.eps}} \caption[np]{The np capture rate calculated in chiral perturbation theory. The predicted capture rate (upper plot) $\sigma=(334\pm 3)$ mb agrees with the experimental value $(334.2\pm 0.5)$ mb. The lower plot shows the ratios of the matrix elements of the two-body terms over the one-body term. ``Tree" corresponds to the leading chiral order one-pion exchange term (with the blob in Figs.\ref{graphs}(a,b) replaced by a bare vertex), ``$1\pi (\omega)$" and ``$1\pi (\Delta)$" correspond to the next-to-leading order corrections involving the $\omega$ meson and $\Delta$ resonance, respectively, to the one-pion exchange tree terms Figs.\ref{graphs}(a,b). ``$2\pi$" is the genuine loop correction to the tree contribution. The hadrons appearing in this calculation all have free-space properties. }\label{nprates} \end{figure} A recent calculation \cite{pmr} of the process (\ref{npcap}) is given in Fig.\ref{nprates}. The calculation was done to order $Q^3$ which corresponds to next-to-next-to leading order in chiral expansion, that is, to one-loop order. Given the nucleon mass $m_N$, the pion decay constant $f_\pi$, the pion mass $m_\pi$, the axial coupling constant $g_A$ and the vector meson mass $m_V$ all {\it determined in free space}, all the parameters that appear in the theory are fixed in the theory except for the hard core radius $r_c$ in the wave function reflecting on our inability to handle in ChPT very short-range physics. The remarkable agreement with the experiment shows that the two-nucleon systems we are looking at are made up of two nucleons with their properties as given in free space: the chiral Lagrangian with the vacuum values of hadron parameters describes nature remarkably well at low chiral orders. \begin{figure} \centerline{\epsfig{file=pmr3fig1.eps}} \caption{Generic graphs contributing to exchange currents. (a) and (b) are one-pion exchange and (c) represents multipion and/or heavy-meson exchange currents. The large filled circles represent one-nucleon, one-pion irreducible graphs, the solid line the nucleon, the dotted line the pion and the wiggly line the current $j_\mu$.}\label{graphs} \end{figure} A very valuable information for heavier and denser nuclei is lodged in some of the terms that are {\it negligible} in the process (\ref{npcap}). As I showed elsewhere \cite{mr91}, two-body currents involving four baryon fields (e.g. Fig.\ref{graphs}(c) with the blob replaced by a point) like \begin{eqnarray} \bar{B}\Gamma_\mu B \bar{B}\tilde{\Gamma} B\label{fourbaryon} \end{eqnarray} with $\tilde{\Gamma}$ and $\Gamma_\mu$ representing some Lorentz scalar and Lorentz vector quantities consistent with chiral symmetry, respectively, are subdominant in the chiral counting and can be ignored. Now terms like (\ref{fourbaryon}) show up in the chiral Lagrangian as a result of integrating out heavy degrees of freedom with an energy scale $E\roughly> \Lambda_\chi$. Consider for instance the scalar meson $\sigma$ that plays an important role in effective field theory of nuclei, a prototype of which being the Walecka model. In free space, there is no low-lying scalar that can appear in the low-energy chiral Lagrangian. But there is a high-lying scalar field that can be associated with the trace anomaly of QCD \begin{eqnarray} (T_\mu^\mu)_{QCD}=-\frac{\beta (g)}{2g} (G_{\mu\nu}^a)^2\sim \chi^4 \end{eqnarray} where $\beta$ is the QCD beta function, $g$ the color gauge coupling constant, $G_{\mu\nu}^a$ the gluon field tensor and $\chi$ the scalar glueball field. Now the $\chi$ field is massive, with $m_\chi\sim 2$ GeV, so this degree of freedom appears only in the counter terms. It will give rise to a term like \begin{eqnarray} \kappa \bar{B}j_\mu^{(1)}B \bar{B}B\label{sigmacurrent} \end{eqnarray} with the coefficient $\kappa$ suppressed by the power $(Q/m_\chi)^2$. In the process (\ref{npcap}), $Q$ is of order of 40 MeV, so $(Q/m_\chi)^2 < 10^{-3}$. Stated differently an effective two-body term like (\ref{sigmacurrent}) will be screened by the short-range correlation implicit in the wave functions. The situation is quite different, however, in dense nuclear medium. As discussed in \cite{brpr}, as density increases, the scalar (or precisely the quarkish component of the scalar) $\chi$ moves downwards in energy and at some high density, it joins the triplet of nearly massless pions to make up the quartet of the $O(4)$ symmetry of chiral symmetry. The merging presumably takes place at the chiral transition point discussed below. The point is that as discussed by Beane and van Kolck \cite{beane}, in order to reconcile Weinberg's ``mended symmetry" \cite{mendedwein} with effective chiral Lagrangians at some shorter length scale, the scalar field must come down as a dilaton. Now if it comes down below the chiral scale $\Lambda_\chi$, then we can no longer consider the scalar as a counter-term contribution. It will strongly couple to low-mass multipion excitations giving among others what is usually taken as a scalar field $\sigma$ in effective nuclear forces \cite{brpr}. All other hadrons (other than Goldstone bosons) will also couple to this scalar as well and will undergo a mass shift as density increases. A chiral Lagrangian that accounts for this phenomenon has been shown to lead to the Brown-Rho scaling \cite{br91} \begin{eqnarray} m_B^\star/m_B\approx m_V^\star/m_V\approx m_\sigma^\star/m_\sigma \approx f_\pi^\star/f_\pi\approx \cdots\label{BR} \end{eqnarray} where $B$ stands for baryons, $V$ for vector mesons and $\sigma$ for the dilatonic scalar. The star stands for density-dependent quantities. Such a chiral Lagrangian effective in dense system will then contain these effective constants instead of the free-space values used in the process (\ref{npcap}) while preserving the free-space chiral symmetry. As I will argue, the consequence of this scaling can be significant in heavy nuclei and nuclear matter. One can already see the effect of this scaling in finite nuclei. One clear case is the axial charge transition in nuclei. It was shown in \cite{pmr,pmr93} that to order $Q^3$ in chiral expansion the axial charge transition matrix element in heavy nuclei \begin{eqnarray} A (0^\pm)\rightarrow B(0^\mp),\ \ \ \ \Delta T=1\label{axial} \end{eqnarray} is enhanced with respect to the impulse approximation by the factor \begin{eqnarray} \epsilon_{\tiny MEC}= \frac{m_N}{m_N^\star} (1+ R) \end{eqnarray} where $R$ is the ratio of the exchange current matrix element Fig.\ref{graphs} to the impulse approximation calculated with matter-free-space constants. In (\ref{axial}), $R$ is essentially given by Fig.\ref{graphs}(a) with the bare coupling and is given by $R\approx 0.5$ with a small variation with density of the system. Now at nuclear matter density, $m_N^\star/m_N\approx 0.75$, so \begin{eqnarray} \epsilon_{\tiny MEC}\approx 2. \end{eqnarray} In light nuclei, we expect $\epsilon_{\tiny MEC}\roughly> 1.5$. These results are in agreement what was found experimentally \cite{warburton}. One predicts a similar effect in magnetic moments of heavy nuclei but here one has to include other effects of equal importance present in the vector current case. For instance, the ``back-flow" correction due to Galilean invariance cancels almost completely the corrections coming from the scaled nucleon mass. We now turn to an important issue of making a bridge between the chiral theory and Walecka mean field theory of nuclei and nuclear matter which is found to be very successful. I wish to show here, following \cite{br95-1}, that Walecka theory is equivalent to the chiral Lagrangian theory at mean field {\it with} the BR scaling and that this would allow a treatment of fluctuations into different flavor directions (such as strangeness) in a way consistent with the properties of normal nuclear matter. To do this we can focus on the four-fermi interactions allowed in chiral Lagrangians that are relevant in making contact with Walecka theory, \begin{eqnarray} {\cal L}_{4f}=\alpha \left(\bar{B}B\right)^2 +\beta \left(\bar{B} v^\mu B\right)^2 \label{4flag} \end{eqnarray} where $\alpha$ and $\beta$ are dimension -2 constants and we are using the heavy-fermion formalism so that $v_\mu$ is the velocity four-vector of the heavy baryon. Let us imagine that the first term of (\ref{4flag}) arises from integrating out the heavy chiral singlet scalar $\chi$ and the second term from integrating out a heavy chiral singlet vector meson $\omega$. We can include other degrees of freedom in a similar way but we will not need them for symmetric nuclear matter that we shall consider. In this case, we can identify the constants \begin{eqnarray} \alpha=\frac{g_\chi^2}{2m_\chi^2}, \ \ \ \ \beta=-\frac{g_\omega^2}{2m_\omega^2}. \end{eqnarray} We now ask what happens to this Lagrangian when it is immersed in dense and/or hot matter. In mean field, we get the nucleon scalar potential $S_N$ and vector potential $V_N$ as \begin{eqnarray} S_N &=& -\frac{{g_\sigma^\star}^2}{{m_\sigma^\star}^2} \rho_s,\label{SN}\\ V_N &=& \frac{9}{{8f_\pi^\star}^2} \rho\label{VN} \end{eqnarray} where $\rho_s$ is the scalar density and $\rho$ the vector density. In obtaining (\ref{VN}), we have used $SU(3)_f$ relations together with KSRF relation which is known to hold well and put the stars in (\ref{SN}) and (\ref{VN}) to indicate that they are in-medium quantities. Now comparing with the phenomenology with Walecka model, we find that the identification requires that (\ref{BR}) holds with \begin{eqnarray} \frac{f_\pi^\star}{f_\pi}\approx 0.77. \end{eqnarray} Together with what we found in the case of the axial charge transition, we come to the conclusion that the mean field Walecka theory is just the mean field chiral Lagrangian theory {\it with BR scaling}. It is a well-known defect of the mean field Walecka model that the compressibility modulus $K_0$ is much too high in the model. Now how does this defect get rectified? The answer must lie in higher loop corrections going beyond the mean field as the scaling is known to fail to give the nuclear matter saturation \cite{tjon}. This is also seen in recent work of Furnstahl et al \cite{tang} who note that by giving an anomalous dimension 2.7 to the scalar field $\sigma$ with the Lagrangian suitably implemented with the trace anomaly, they can obtain the low $K_0\approx 200$ MeV and the suppression of the many-body terms $\sigma^n$, $n >2$. It is plausible that the anomalous dimension is mocking up the quantum loop effects that seem to be needed in the mean field approach given in \cite{br95-1}. A remarkable observation is that at the anomalous dimension of $d_a\approx 2.7$, two things happen simultaneously. One is that the $K_0$ which is large at smaller anomalous dimensions stabilizes at $\sim 200$ MeV for $d_a\approx 2.7$, stays at that value for higher $d_a$'s and secondly, it is at this fine-tuned value of $d_a$ that {\it all} multi-body forces get suppressed. This clearly calls for a simple explanation\cite{chaejun}. The above result immediately suggests how to calculate kaon-nuclear interactions in consistency with the nuclear matter properties as given by Walecka theory. To see this, consider a part of the chiral Lagrangian that figures importantly in the kaon-nuclear sector \begin{eqnarray} {\cal L}_{KN}=\frac{-6i}{8f^2}(\overline{B}\gamma_0 B)\overline{K}\partial_t K + \frac{\Sigma_{KN}}{f^2}(\overline{B}B)\overline{K}K\equiv {\cal L}_\omega +{\cal L}_\sigma\label{kaonL} \end{eqnarray} where $K^T=(K^+ K^0)$. The constant $f$ in (\ref{kaonL}) can be identified in free space with the pion decay constant $f_\pi$. In medium, however, it can be modified as we shall see shortly. In chiral perturbation expansion, the first term corresponds to ${\cal O} (Q)$ and the second term to ${\cal O} (Q^2)$. There is one more ${\cal O} (Q^2)$ term proportional to $\partial_t^2$ which will be taken into account in the numerical results quoted below but they are not important except for quantitative details. One can interpret the first term of (\ref{kaonL}) as arising from integrating out the $\omega$ meson as in the baryon sector. The resulting $K^- N$ vector potential in medium can then be deduced in the same way as for $V_N$: \begin{eqnarray} V_{K^\pm}=\pm\frac{3}{8{f^\star_\pi}^2}\rho. \end{eqnarray} Thus in medium, we may set $f\approx f_\pi^\star$ and obtain \begin{eqnarray} V_{K^\pm}=\pm\frac 13 V_N.\label{omegascale} \end{eqnarray} This just says that the $\omega$ couples to a {\it matter field} kaon, hence 1/3 of the $\omega$ coupling to the nucleon. The reason for this matter-field nature of the kaon is that all nonstrange hadrons become light in dense medium, so the kaon becomes in some sense heavy. This dual character is known from the hyperon structure which is well described by considering the kaon to be heavy as in the Callan-Klebanov model. As for the second term of (\ref{kaonL}), we use that the kaon behaves as a massive matter field. We therefore expect that it be coupled to the chiral scalar $\chi$ as \begin{eqnarray} {\cal L}_\sigma = \frac 13 2 m_K g^\star_\sigma \overline{K}K\chi \end{eqnarray} where the factor 1/3 accounts for one non-strange quark in the kaon as compared with three in the nucleon. When the $\chi$ field is integrated out as above, we will get, analogously to the nucleon case, \begin{eqnarray} {\cal L}_\sigma = 2m_K \frac 13 \frac{{g^\star_\sigma}^2}{{m^\star_\sigma}^2}\overline{B}B \overline{K}K. \end{eqnarray} Comparing with the second term of (\ref{kaonL}), we find \begin{eqnarray} \frac{\Sigma_{KN}}{f^2}\approx 2\frac{m_K}{3} \frac{{g^\star_\sigma}^2}{{m^\star_\sigma}^2}.\label{relation} \end{eqnarray} We can get the $\Sigma_{KN}$ from lattice calculations \cite{fukugita}, $\Sigma_{KN}\approx 3.2 m_\pi$. This gives $f\approx f_\pi^\star$. Therefore we have \begin{eqnarray} S_{K^\pm}=\frac 13 S_N. \end{eqnarray} {\it To summarize: the kaon-nuclear potential gotten from a chiral Lagrangian and the nucleon-nuclear potential given by Walecka mean field theory are directly related through BR scaling.} Given Walecka mean fields for nucleons, we can now calculate the corresponding mean-field potential for $K^-$-nuclear interactions in symmetric nuclear matter. {}From the results obtained above, we have \begin{eqnarray} S_{K^-} +V_{K^-}\approx \frac 13 (S_N-V_N). \end{eqnarray} Phenomenology in Walecka mean-field theory gives $(S_N-V_N)\roughly< -600\ {\mbox{MeV}}$ for $\rho=\rho_0$. This leads to the prediction that at nuclear matter density \begin{eqnarray} S_{K^-}+V_{K^-}\roughly< -200\ {\mbox{MeV}}. \end{eqnarray} This seems to be consistent with the result of the analysis in K-mesic atoms made by Friedman, Gal and Batty \cite{friedman} who find attraction at $\rho\approx 0.97\rho_0$ of \begin{eqnarray} S_{K^-}+V_{K^-}=-200\pm 20\ {\mbox{MeV}}. \end{eqnarray} An immediate consequence of this mean field description of the kaonic sector is that kaons will condense in dense neutron star (or nuclear star) matter at a density \begin{eqnarray} \rho_c\sim 2 \rho \end{eqnarray} as found by Lee et al \cite{LBMR} in ChPT to one-loop order. \begin{figure} \vskip -7cm \centerline{\epsfig{file=ceres2.eps}} \vskip -7cm \caption[ceres]{The Li-Ko-Brown explanation of the dilepton data of the CERES collaboration. The dotted line is the theoretical prediction {\it without} the scaling of the $\rho$ and $\omega$ mesons and the solid curve {\it with} the BR scaling. The three lowest mass points are essentially given by Dalitz pairs, so the relevant data points are the ones for higher invariant masses. Note that the peak shown at $M\sim 800$ MeV is predominantly given by the $\omega$ decay outside of the medium.}\label{ceres} \end{figure} To conclude, I make a few remarks on the nature of the scaling properties. \begin{enumerate} \item There are two points to the issue. One is that {\it some or all} light hadrons may be undergoing a downward mass shift as density or temperature is increased. The second is that the scaling is like (\ref{BR}). These are basically two different issues. To the extent that hadron masses are generated spontaneously, it is inevitable that at least some masses should drop. Indeed QCD sum rule calculations do predict the drop for the vector-meson mass \cite{qcdsum}, the most recent value being \cite{jin} \begin{eqnarray} m_\rho^\star/m_\rho=0.78\pm 0.08,\ \ \ \rho\approx \rho_0. \end{eqnarray} On the other hand, the ``universal scaling" (\ref{BR}), obtained at mean field, may not be strictly valid. In fact, large $N_c$ arguments imply $m_N^\star/m_N\approx \sqrt{g_A^\star/g_A} (f_\pi^\star/f_\pi)$. If we look at the strict $N_c=\infty$ limit, $g_A^\star=g_A$ as we showed in \cite{br91}, the scaling (\ref{BR}) holds but we know that in nuclei, $g_A^\star\approx 1$ and thus in finite nuclei we expect that the nucleon scales somewhat faster than the pion decay constant, at least up to $\rho\sim \rho_0$. Whether the ``universal scaling" (\ref{BR}) is consistent with nature remains to be seen. \item An intriguing question is how far the scaling (\ref{BR}) can be pushed in density and/or temperature. Can one use the mean field argument all the way to the chiral phase transition? This is a highly relevant question since there are arguments \cite{kocic} that the second order chiral phase transition relevant to QCD with 2 flavors is of mean field type as in 3D Gross-Neveu model. This suggests that the mean field chiral theory or Walecka theory could be used to discuss the QCD chiral phase transition \cite{br95-2}. If correct, this theory will be very useful for studying phase transitions in heavy ion collisions. \item There are a large number of experimental projects to measure the mass shift of hadrons in dense matter, particularly at GSI, CEBAF and CERN using dilepton-pair production. Recent data from the CERES collaboration \cite{CERES} on $e^+ e^-$ pairs in S on Au collisions at 200 GeV/nucleon can be understood in terms of the scaled mass of the vector mesons in medium, primarily due to density effect. In a recent paper, Li, Ko and Brown \cite{likobrown} have shown that the enhancement of the produced lepton pairs observed in the range of invariant mass $300\ {\mbox{MeV}}\roughly< M \roughly< 550\ {\mbox{MeV}}$ can be explained simply by the BR scaling in the $\rho$ mass as the pairs are produced mainly through $\pi^+ \pi^-\rightarrow \rho^\star \rightarrow e^+ e^-$. The fit is given in Fig.\ref{ceres}. The analysis is a complex one involving the assumption of an expanding fireball in chemical equilibrium, but the economy of the explanation and the quantitative success make it quite compelling. \end{enumerate} \subsection*{Acknowledgments} \indent I am very grateful for continuing discussions with Gerry Brown with whom most of the ideas described here have been developed. I would also like to thank my young collaborators Chang-Hwan Lee, Kurt Langfeld and Tae-Sun Park for their valuable help. This paper was completed at the Center for Theoretical Physics (CTP) of Seoul National University. I would like to thank Dong-Pil Min and the members of the CTP for hospitality and support.
\section{Introduction} In collisions between two heavy nuclei at bombarding energies from a few hundred MeV up to several GeV per nucleon, hadronic matter at high density and temperature is formed. In such collisions a large number of energetic particles are produced and may be used as probes of the hot and dense phase of the reaction \cite{Metag,Cassing,Mosel}. Microscopic transport models, such as BUU and QMD \cite{Cassing,Mosel,Wolf,Aichelin} have been fairly successful in describing particle production in heavy-ion reactions. In these transport models, the nucleons propagate in an effective one-body field while subject to direct two-body collisions. Sufficiently energetic nucleon-nucleon collisions may agitate one or both of the colliding nucleons to a nucleon resonance, especially $\Delta(1232)$, $N^*(1440)$, and $N^*(1535)$. Such resonances propagate in their own mean field and may collide with nucleons or other nucleon resonances as well. Furthermore, the nucleon resonances may decay by meson emission and these decay processes constitute the main mechanisms for the production of energetic mesons \cite{Mosel}. The transport descriptions normally employ the vacuum properties of the resonances and mesons, \ie\ the needed cross sections, decay widths, and dispersion relations are taken according to their values in vacuum \cite{Wolf}. However, in infinite nuclear matter, a system of interacting $\pi$ mesons, nucleons, and $\Delta$ isobars will couple to form spin-isospin modes. Some of these modes are non-collective in their character, dominated by a single baryon-hole excitation, while other modes are collective and correspond to meson-like states (quasimesons). The non-collective spin-isospin modes correspond to those already included in transport simulations by promoting a nucleon from below to above the Fermi surface. The collective spin-isospin modes can effectively be regarded as particles of mesonic character (quasimesons), which by means of a local density and temperature approximation can be incorporated into the transport descriptions. Some in-medium modifications have already been employed in calculations of heavy-ion collisions, both qualitatively \cite{Weise} and by transport simulations \cite{Bertsch,Giessen,Texas}. A more elaborate $\pi + N N^{-1} + \Delta N^{-1}$ model was employed in ref.\ \cite{main} to derive several quantities useful for implementation of in-medium properties in transport descriptions. While it is straightforward to apply the local density approximation in the interior regions of the nuclear system, there are conceptual problems of how to proceed at the nuclear surface where the density approaches zero and the quasimesons convert to real physical particles. The problem is that while in a real system no hole states exist in vacuum, a collective ($\Delta N^{-1}$-like) mode can in a stationary infinite system exist for an arbitrary small (but finite) density. This is because the particles in the stationary system have infinite time to explore the entire system and form collective modes also at extremely low densities. In this paper we will therefore discuss how a proper conversion from quasimesons to real particles at the surface can be performed. Earlier works \cite{Giessen,Texas} have treated this conversion by various approximations (see further the discussion in section \ref{sec_Qpions}). In this work we will present a somewhat different approach, based on ref.\ \cite{main}, and discuss how the approximations in the earlier works can be improved upon. The present paper constitutes a qualitative investigation of treatment of pionic modes at a nuclear surface in transport simulations. To make the presentation simple and transparent we will therefore restrict ourselves to some special cases of the more complete investigation in ref.\ \cite{main}. We will only consider the collective modes in the spin-longitudinal channel (pion like), since this is the dominant channel at the energies we have in mind in this paper. Collective modes in the spin-transverse channel ($\rho$ meson like) can be treated completely analogously. Furthermore we will consider the zero-temperature case, and the $\Delta$ width will not be included in the calculation of the dispersion relations (denoted as the reference case in ref.\ \cite{main}). The justification for considering only $T=0$ in this paper is that there is not a strong dependence on the dispersion relations of the collective modes for moderately large temperatures, especially not at low densities at the surface. Also, the temperature is not expected to be very high at the surface. However we want to emphasize that there is no principal difficulty associated with incorporating $T>0$ in the treatment. The motivation for omitting the $\Delta$ width, in the calculation of the dispersion relations, is somewhat more involved. Including the $\Delta$ width self-consistently in the calculation of the spin-isospin modes encompasses decay processes like \[ \tilde{\pi}_j \rightarrow \Delta N^{-1} \rightarrow (N+\tilde{\pi}_k) N^{-1}\ . \] However, since such processes are already explicitly contained in the transport simulation by processes like \[ \tilde{\pi}_j + N \rightarrow \Delta \rightarrow N+\tilde{\pi}_k \ , \] it does not seem to be correct to include the entire self-consistent $\Delta$ width when calculating the collective modes to be used in the transport description. Instead it seems more correct to use the results obtained with the $\Delta$ width omitted, both for the energies of the modes and for the partial $\Delta$ widths to be used in the decays $\Delta \rightarrow N +\tilde{\pi}_j$. However one should note that by omitting the $\Delta$ width in the dispersion relations one also fails to take into account the fact that the pionic modes have a Breit-Wigner like energy distribution, analogous with the $\Delta$. The width and center of this distribution are determined by the the self-consistent $\Delta$ width and depends on the particular pionic mode and its momentum. The center of the distribution approximately corresponds to the energy found when the $\Delta$ width is omitted \cite{main}. In section \ref{sec_Model} we will give a brief presentation of the model. The dispersion relations and amplitudes of the spin-isospin modes obtained in infinite nuclear matter, are presented and discussed in section \ref{sec_DispAmp}. Section \ref{sec_Qpions} is devoted to a discussion of the pionic modes at a nuclear surface and the implications for transport descriptions, while our results are summarized in section \ref{sec_Sum}. In addition we present in appendix \ref{sec_Refl} a discussion of reflection and transmission properties at the nuclear surface, and in appendix \ref{sec_RPAsolu} some technical details for the RPA equations. \section{The model} \label{sec_Model} The model presented in this section is treated and motivated in detail in ref.\ \cite{main}. For convenience we here present a brief recapitulation of the essential points. Furthermore, the presentation in this section only treats the spin-longitudinal channel for the special case when $T=0$ and the $\Delta$ width is omitted in the calculation of the spin-isospin modes. \subsection{Spin-isospin modes in an infinite system} \label{sec_Model-1} We consider a system of interacting nucleons ($N$), delta isobars ($\Delta$) and pi mesons ($\pi$). In order to investigate the in-medium properties of the interacting particles, we employ a cubic box with side length $L$; the calculated properties are not sensitive to the actual size, so we need not take the limit $L\to\infty$ explicitly. The in-medium properties are obtained by using the Green's function technique, starting from non-interacting hadrons. The non-interacting Hamiltonian can be written \begin{equation} H_0 = \sum_k e_k \hat{b}^{\dagger}_k \hat{b}^{ }_k + \sum_l \hbar \omega_\pi(\mathbf{q_l}) \hat{\pi}^{\dagger}_l \hat{\pi}^{ }_l\ . \label{eq_H0} \end{equation} Here the index $k=(\mathbf{p}_k; \, s_k, m_{s_k}; \, t_k, m_{t_k})$ represents the baryon momentum, spin, and isospin. The spin and isospin quantum numbers, $s_k$ and $t_k$, take the values $1\over2$ and $3\over2$ for $N$ and $\Delta$, respectively. The energy of baryon $k$ moving in a (spatially constant) potential is denoted $e_k$. The baryon creation and annihilation operators, $\hat{b}^{\dagger}_k$ and $\hat{b}^{ }_k$, are normalized such that they satisfy the usual anti-commutation relation, \begin{equation} \{ \hat{b}^{\dagger}_k, \: \hat{b}^{\mbox{ }}_{k'} \} = \delta_{k,k'}\ . \end{equation} In the pion part of $H_0$, the index $l$ represents the pion momentum and isospin, $l = (\mathbf{p}_l,\lambda_l=0,\pm~1)$. The meson energy is given by $\hbar \omega_{\pi} = [m_{\pi}^2 + \mathbf{q}^2]^{1/2}$ and the creation and annihilation operators of the pion are normalized such that they satisfy the usual commutation relation, \begin{equation} [ \hat{\pi}^{\dagger}_l, \: \hat{\pi}^{ }_{l'} ] = \delta_{l,l'}\ . \end{equation} Note that the $\Delta$ isobar described by $H_0$ has no decay width, $\Gamma_\Delta=0$. When the interactions are turned on, the $\Delta$ width will emerge and it will then automatically include also the free width. \subsubsection{Basic interactions} At the $N \pi N$ and $N \pi \Delta$ vertices we will use effective $p$-wave interactions, $V_{N \pi N}$ and $V_{N \pi \Delta}$, which in the momentum representation can be written as \cite{main,OTW} \begin{eqnarray} &~& V_{N \pi N} = i c \: \frac{(\hbar c)^{1\over2}}{L^3} \: \left[ \frac{2m_N c^2}{m_N c^2 + \sqrt{s}} \right]^{1\over2} \frac{f^\pi_{NN}}{m_\pi c^2}\ F_{\pi}(q) \; (\mathbf{\sigma} \mathbf{\cdot} \mathbf{q}_{cm}) \: \veC{\tau} \mathbf{\cdot} \Vec{\phi}_\pi(\mathbf{q}) \label{eq_Vnpn} \\ &~& V_{N \pi \Delta} = i c \: \frac{(\hbar c)^{1\over2}}{L^3} \: \left[ \frac{2m_\Delta c^2}{m_\Delta c^2 + \sqrt{s}}\right]^{1\over2} \frac{f^\pi_{N\Delta}}{m_\pi c^2}\ F_{\pi}(q) \; (\mathbf{S^+} \mathbf{\cdot} \mathbf{q}_{cm}) \: \veC{T}^+ \mathbf{\cdot} \Vec{\phi}_\pi(\mathbf{q}) + {\rm h.c.} \phantom{123} \label{eq_Vnpd} \end{eqnarray} In these expressions, $\sqrt{s}$ is the center-of-mass energy in the $N \pi$ system and $\mathbf{q}_{cm}$ is the pion momentum in the $N \pi$ center-of-mass system, which in the non-relativistic limit is given by \begin{equation} \mathbf{q}_{cm} \approx \frac{m_N c^2}{m_N c^2 + \hbar \omega}\ \mathbf{q} - \frac{\hbar \omega} {m_N c^2 + \hbar \omega}\ \mathbf{p}_N\ , \label{eq_qcm} \end{equation} where $\hbar \omega$ and $\mathbf{q}$ is the pion energy and momentum, and $\mathbf{p}_N$ is the nucleon momentum in an arbitrary frame. The Pauli spin and isospin matrices are denoted $\mathbf{\sigma}$ and $\vec{\tau}$, and $\mathbf{S^+}$ and $\vec{T}^+$ are spin and isospin $1\over2$ to $3\over2$ transition operators normalized such that $<{3\over2},{3\over2}|S^+_{+1}|{1\over2},{1\over2}>=1$.\footnote{For clarity, we generally employ bold-face characters to denote quantities with vector and tensor properties under ordinary spatial rotations, while arrows are employed to indicate the transformation properties under rotations in isospace.} The momentum representation of the pion field is taken as \begin{equation} \phi^\pi_\lambda(\mathbf{q}) = \frac{L^{3/2} \hbar c}{\sqrt{2 \hbar \omega_\pi(\mathbf{q})}} \left[ \hat{\pi}^{ }_\lambda(\mathbf{q}) + (-1)^\lambda \hat{\pi}_{-\lambda}^\dagger(-\mathbf{q}) \right]\ . \end{equation} The interactions contain a monopole form factor, \begin{equation} F_{\pi}(q) = \frac{\Lambda_\pi^2 - (m_\pi c^2)^2}{\Lambda_\pi^2 - (cq)^2}\ , \label{eq_Fpi} \end{equation} and the coupling constants are determined at $(cq)^2 = (\hbar \omega)^2 - (c \, \mathbf{q})^2 = (m_\pi c^2 )^2$ and $\sqrt{s} = m_N c^2$ or $\sqrt{s} = m_\Delta c^2$. In addition we will also include effective short-range interactions at nucleon-hole vertices, again written in momentum space, \begin{equation} V_{NN^{-1},NN^{-1}} = \: \left( \frac{\hbar c}{L} \right)^3 \: g_{NN}' \left( \frac{f^\pi_{NN}}{m_\pi} \right)^2 |F_g(q)|^2\ (\mathbf{\sigma_1 \cdot \sigma_2}) (\veC{\tau_1} \cdot \Vec{\tau_2})\ , \label{eq_Vg} \end{equation} and the corresponding interactions obtained when one (or two) of the nucleons is replaced by a $\Delta$. The strength of the short-range interactions is determined by the correlation parameters $g_{NN}'$, $g_{N\Delta}'$, and $g_{\Delta \Delta}'$. \subsection{RPA approximation} We want to calculate a spin-isospin mode Green's function within the RPA approximation, symbolically \begin{equation} G^{\rm RPA}(\alpha,\beta;\omega) = G_{0}(\alpha,\beta;\omega) + \sum_{\gamma,\kappa} G_{0}(\alpha,\gamma;\omega)\ {\cal V}(\gamma,\kappa;\omega)\ G^{\rm RPA}(\kappa,\beta;\omega)\ . \label{eq_GRPA} \end{equation} The spin-isospin modes, here represented by the Green's function $G^{\rm RPA}$, will in this approximation be obtained as an infinite iteration of (non-interacting) pion, nucleon-hole, and $\Delta$-hole states, represented by the diagonal Green's function $G_0$, coupled with the interactions specified in eqs.\ (\ref{eq_Vnpn}--\ref{eq_Vg}) which here are summarized by the symbolic interaction ${\cal V}$. In nuclear collisions at beam energies up to about one GeV per nucleon, which is the domain of application that we have in mind, only relatively few mesons and isobars are produced and so the associated quantum-statistical effects may be ignored. Accordingly, we assume $n_\Delta \ll 1$, and $n_\pi \ll 1$. A set of RPA equations, equivalent to eq.\ (\ref{eq_GRPA}) were derived in ref.\ \cite{main}. {}From these equations eigenvectors and eigenenergies are obtained for the different spin-isospin modes. The eigenvectors will yield the amplitudes of the different components ($\pi$, $NN^{-1}$, $\Delta N^{-1}$) forming the particular spin-isospin mode with the given eigenenergy. These RPA amplitudes contain important information about the nature of the different spin-isospin modes. The spin-isospin modes (or excited RPA states), $|\Psi_\nu>$, are created by an operator $Q^{\dagger}_\nu$, \begin{equation} Q^{\dagger}_\nu(\mathbf{q},\lambda) = \sum_{jk} X^\nu_{jk}(\mathbf{q},\lambda) \hat{b}^{\dagger}_j \hat{b}^{ }_k + \sum_{k} Z^\nu_k(\mathbf{q},\lambda) \, \hat{\pi}_k^\dagger - \sum_{k} W^\nu_k(\mathbf{q},\lambda) \, \hat{\pi}^{ }_k\ . \label{eq_Qrpal} \end{equation} The quantity $X^\nu_{jk}(\mathbf{q},\lambda)$ is here the amplitude of the baryon-hole ($N N^{-1}$ or $\Delta N^{-1}$) component of the spin-isospin mode $|\Psi_\nu>$ at momentum $\mathbf{q}$ and isospin $\lambda$, while $Z^\nu_k(\mathbf{q},\lambda)$ and $W^\nu_k(\mathbf{q},\lambda)$, in the same way, are the amplitudes of the pionic component. The summation over baryon and meson states in eq.\ (\ref{eq_Qrpal}) is restricted by taking $X_{jk} \propto \delta_{\mathbf{p}_j, \mathbf{p}_k + \mathbf{q}}$, $Z_k \propto \delta_{\mathbf{p}_k, \mathbf{q}} \delta_{\lambda_k, \lambda}$, and $W_k \propto \delta_{\mathbf{p}_k, -\mathbf{q}} \delta_{\lambda_k, -\lambda}$. The RPA equations are obtained from the relation \begin{equation} <[\delta Q,[H,Q^{\dagger}]]> = \hbar \omega <[\delta Q,Q^{\dagger}]>\ , \label{eq_RPA-gen} \end{equation} with $\delta Q=\hat{b}^{\dagger}_k\hat{b}^{ }_j$, $\hat{\pi}_r$, or $\hat{\pi}_r^\dagger$, and where the brackets $<\cdot>$ denote the expectation value in the interacting ground state. It can be shown that the set of RPA solutions constitutes an orthonormal set. For convenience the solutions of the RPA equations of ref.\ \cite{main}, are recapitulated in appendix \ref{sec_RPAsolu}. \subsubsection{The total $\Delta$ width} The $\Delta$ self energy $\Sigma_\Delta$ is calculated according to the diagrams in fig.\ \ref{fig_DseGraph}, by taking into account all the diagrams corresponding to the $\Delta$ decaying into a spin-isospin mode and a nucleon, which then again form a $\Delta$. In the spin-longitudinal channel we obtain (ref.\ \cite{main}) \begin{equation} \Gamma^l_\Delta(E_\Delta,\mathbf{p}_\Delta) = \mbox{Im } \frac{2}{3} \left( \frac{\hbar c}{L} \right)^3 \sum_{\mathbf{q}} \, \left[ \theta({\cal E}) - n(\mathbf{p}_\Delta-\mathbf{q}) \right] \bar{M}(\Delta N,N \Delta)\ . \label{eq_DSE2} \end{equation} where the energy available for the spin-isospin mode is given by \begin{equation} {\cal E} = E_\Delta - e_N(\mathbf{p}_\Delta-\mathbf{q})\ , \end{equation} and $\bar{M}(34,12)$ can be expressed as \begin{equation} \bar{M}(34,12) = \sum_{\omega_\nu > 0} \left\{ \frac{ h(31;\nu) h(24;\nu) } { \hbar \omega - \hbar \omega_\nu + i \eta } - \frac{ h(31;\nu) h(24;\nu) } { \hbar \omega + \hbar \omega_\nu - i \eta } \right\} + \frac{ f^\pi_{31} f^\pi_{24} }{(m_\pi c^2)^2} F_g^2 g'_{34,12} \ . \label{eq_Mrpa} \end{equation} The factor $h(jk,\nu)$ is obtained from the interactions at the vertex consisting of baryons $j$ and $k$, and the spin-isospin mode $\nu$. The interactions to be used depend on the non-interacting states that the mode consists of and must therefore be multiplied by the amplitude of the corresponding state, \begin{equation} h(jk,\nu) \vartheta^l(jk) = \frac{V_{j \pi k}}{\sqrt{2 \hbar \omega_\pi}} \, [Z(\nu)+W(\nu)] + \sum_{mn} V_{jk,mn} \, X_{mn}(\nu) \ , \label{eq_hMotiv} \end{equation} where $\vartheta^l(jk)$ is a short hand notation for the spin-isospin matrix elements in the spin-longitudinal channel, $V_{j \pi k}$ is defined in (\ref{eq_Vnpn}) and (\ref{eq_Vnpd}), $V_{jk,mn}$ is defined in (\ref{eq_Vg}), and the amplitudes $X^l_{mn}$, $Z^l$, $W^l$ are defined in eq.\ (\ref{eq_Qrpal}). The explicit expression for $h(jk,\nu)$ is somewhat lengthy and has therefore also been relegated to appendix \ref{sec_RPAsolu}. \subsubsection{Specific $\Delta$ channels} The total $\Delta$ width gives the transition probability per unit time for the $\Delta$ resonance to decay to any of its decay channels. In a transport description one explicitly allows the $\Delta$ resonance to decay into specific final particles. Consequently, one needs not only the total $\Delta$ width (which is the sum of all decay channels) but also the partial widths governing the decay into specific RPA channels. These decay channels consist of a nucleon and one of the spin-isospin modes. Since we have access to all the amplitudes of a given spin-isospin mode on the different unperturbed states, it is possible to derive an expression for the partial contribution to $\Gamma_\Delta$ from the $\Delta$ decay to a specific mode $\nu$. The right-hand side of fig.\ \ref{fig_DseGraph} shows a diagrammatic representation of such a process. The partial $\Delta$ width for a $\Delta$ decay to a nucleon and a spin-longitudinal mode $\nu$ becomes \cite{main} \begin{eqnarray} \tilde{\Gamma}_\Delta^\nu(E_\Delta,\mathbf{p}_\Delta) & = & \int \frac{d^3p_N}{(2\pi)^3} \, \frac{d^3q}{(2\pi)^3} \; | \frac{V_{\Delta \pi N}}{\sqrt{2 \hbar \omega_\pi}} \cdot [Z(\nu)+W(\nu)] + \sum_{mn} V_{\Delta N,mn} \cdot X_{mn}(\nu)|^2 \nonumber \\ & ~& \times \bar{n}_N(\mathbf{p}_N) \, (2\pi)^3 \delta( \mathbf{p}_\Delta - \mathbf{p}_N - \mathbf{q} ) 2\pi \delta( E_\Delta - e_N - \hbar \omega ) \nonumber \\ & = & \frac{1}{3} \int \frac{d^3q}{(2\pi)^3} \: | h(\Delta N,\omega_\nu) |^2 \, \bar{n}_N(\mathbf{p}_\Delta - \mathbf{q}) \, 2\pi \delta( E_\Delta - e_N - \hbar \omega_\nu ) \label{eq_GnuReal} \end{eqnarray} where the factor, $\bar{n}_N = 1-n_N$, takes into account the Pauli blocking of the nucleon. Note that when this expression is to be used in transport models the factor $\bar{n}_N$ should be omitted since the Pauli blocking is treated explicitly in the transport description. The expression (\ref{eq_GnuReal}) is identical to the contribution from one of the $\nu$ terms in eq.\ (\ref{eq_DSE2}). \section{ Dispersion relations and amplitudes } \label{sec_DispAmp} {}From eq.\ (\ref{eq_Xx}) in appendix \ref{sec_RPAsolu} we calculate the energies of the spin-isospin modes that are formed in the interacting system, \ie\ their dispersion relations. Fig.\ \ref{fig_DispT0r10} displays the dispersion relations at normal nuclear density, $\rho = \rho^0 = 0.153\ \fm^{-3}$. In fig.\ \ref{fig_DispT0r10} a number of different modes in the spin-longitudinal ($\pi$-like) channel are apparent. Some of those are non-collective $NN^{-1}$ modes (solid curves), which have their energies within the regions \begin{eqnarray} 0 & \leq & \hbar \omega\ \leq\ \frac{q^2}{2 m_N^*} + \frac{q p_F}{m_N^*}\ , \qquad q < 2 p_F \ , \nonumber \\ \frac{q^2}{2 m_N^*} - \frac{q p_F}{m_N^*} & \leq & \hbar \omega\ \leq\ \frac{q^2}{2 m_N^*} + \frac{q p_F}{m_N^*}\ , \qquad q > 2 p_F\ . \label{eq_NhCont} \end{eqnarray} Since we are presenting our results for a box normalization with a finite side length $L$, we obtain a discrete number of non-collective $NN^{-1}$ modes. The total number of spin-isospin modes within the region (\ref{eq_NhCont}) depends on $L$ and tends towards a continuum in the limit $L \rightarrow \infty$. Similarly, a number of non-collective $\Delta N^{-1}$ states emerge in fig.\ \ref{fig_DispT0r10} which, for a fixed $\mathbf{q}$, have their energies constrained to a band, \begin{equation} m_\Delta - m_N + \frac{q^2}{2 m_\Delta} - \frac{q p_F}{m_\Delta}\ \leq\ \hbar\omega\ \leq\ m_\Delta - m_N + \frac{q^2}{2 m_\Delta} + \frac{q p_F}{m_\Delta}\ . \label{eq_DhCont} \end{equation} The non-collective baryon-hole modes correspond in a transport description to propagation of uncoupled baryons ($N$ or $\Delta$). This was discussed and studied in detail in ref.\ \cite{main} and will therefore not be further discussed in this paper. In addition, two collective modes appear in fig.\ \ref{fig_DispT0r10}, represented by dot-dashed curves. The lower one starts at $\hbar \omega = m_\pi c^2$ at $q = 0$ and continues into the $\Delta N^{-1}$ region around $q \approx 360\ \MeV/c$. This mode will in the following be referred to as $\tilde{\pi}_1$. The upper collective mode starts slightly above $\hbar\omega \approx m_\Delta c^2 - m_N c^2$ at $q = 0$ and approaches $\hbar\omega_\pi = [(m_\pi c^2)^2 + (cq)^2]^{1/2}$ at large $q$. This mode is denoted $\tilde{\pi}_2$. The incorporation of the two collective spin-isospin modes into transport equations is more involved. These modes can be regarded as separate particles of pionic character, $\tilde{\pi}_1$ and $\tilde{\pi}_2$, and treated in a manner analogous to the standard treatment of the pion. Since the pion is then fully included in the description, it should no longer be treated explicitly. The propagation of the two collective pionic modes is governed by the effective Hamiltonians \begin{eqnarray} \tilde{H}_1(\mathbf{r},\mathbf{q}) & = & \hbar \omega_{1}(\mathbf{q};\rho(\mathbf{r}))\ \equiv\ \hbar \tilde{\omega}_1\ , \nonumber \\ \tilde{H}_2(\mathbf{r},\mathbf{q}) & = & \hbar \omega_{2}(\mathbf{q};\rho(\mathbf{r}))\ \equiv\ \hbar \tilde{\omega}_2\ , \label{eq_Hqpion} \end{eqnarray} where $\hbar \omega_{1}$ and $\hbar \omega_{2}$ are the energy-momentum relations for the lower and upper collective modes displayed in fig.\ \ref{fig_DispT0r10} for $\rho=\rho^0$. Note that the spatial dependence of $\tilde{H}_1(\mathbf{r},\mathbf{q})$ is incorporated by representing $\rho(\mathbf{r})$ as a local quantity. To facilitate center-of-mass transformations we parametrize the dispersion relations of the pionic modes in the form \begin{equation} \hbar \tilde{\omega}(\mathbf{q};\rho) \approx \left\{ [c \, q - c \, q_0(\rho)]^2 + m_0(\rho)^2 c^4 \right\}^{\frac{1}{2}} + U_0(\rho) \label{eq_DispParam} \end{equation} For convenience in the transport simulations, we have chosen to use relatively simple expressions for the parametrization, rather than to try to optimize the fit. In this way a quasipion moves like a relativistic particle with the group velocity determined by an effective momentum and energy \begin{equation} \frac{d \hbar \tilde{\omega}}{dq} = c \frac{c(q-q_0)}{\hbar \tilde{\omega} -U_0} = c \frac{c q^*}{\hbar \tilde{\omega}^*}\ . \end{equation} The density-dependent parameters $q_0$, $m_0$ and $U_0$ are presented in fig.\ \ref{fig_DispParam}. Furthermore, in the collision term of the standard transport description the process for the production and absorption of pions $\Delta \leftrightarrow N + \pi$, should be replaced by the two distinct processes \begin{equation} \Delta \leftrightarrow N + \tilde{\pi}_1 \quad \mbox{ and } \quad \Delta \leftrightarrow N + \tilde{\pi}_2\ . \label{eq_Ddecay} \end{equation} The $\Delta$ decay is governed by the $\Delta$ decay width in the medium to these two specific channels, $\tilde{\Gamma}_\Delta^j$. These partial widths, should be employed in the same manner as the free width, \ie\ they describe the probability for the $\Delta$ isobar to decay into a nucleon and a pion. The only difference is that several collective pionic modes are available in the final state. In fig.\ \ref{fig_GamParNPB} we present total and partial $\Delta$ widths for a $\Delta$ with the momentum 300 MeV/$c$. The reverse processes in (\ref{eq_Ddecay}) are characterized by cross sections that were presented and discussed in ref.\ \cite{main}. To obtain the partial $\Delta$ width from eq.\ (\ref{eq_GnuReal}) it is necessary to know the amplitudes $Z$, $\sum X_{\Delta N^{-1}}$ and $\sum X_{N N^{-1}}$. We therefore also present a parametrization of these quantities. On the lower pionic mode the pionic component dominates for small momenta, $q$, and the $\Delta N^{-1}$ component dominates at larger momenta. Therefore the sum of all individual $\Delta N^{-1}$ components will for small $q$ increase with $q$. However, when the lower pionic mode enters the $\Delta N^{-1}$ region the collectivity disappears gradually, and the sum of all individual $\Delta N^{-1}$ components starts to decrease with $q$. Thus we employ the form \begin{eqnarray} \sum X^{\tilde{\pi}_1}_{\Delta N^{-1}}(\mathbf{q};\rho) & \approx & f_\Delta(q,\, \vec{C}[X^{\tilde{\pi}_1}_{\Delta N^{-1}}])\ , \label{eq_XDh-low-parm} \\ \sum X^{\tilde{\pi}_1}_{N N^{-1}}(\mathbf{q};\rho) & \approx & f_N(q,\, \vec{C}[X^{\tilde{\pi}_1}_{ N N^{-1}}])\ , \label{eq_XNh-low-parm} \end{eqnarray} with \begin{equation} f_{\stackrel{\scriptstyle \Delta}{\scriptstyle N} }(q,\vec{C}) = \frac{ \pm C_1^2 + C_2 \, cq }{ C_3^2 + (c q - C_4)^2 }\ . \label{eq_f1(q,rho)-parm} \end{equation} The density-dependent coefficients $\vec{C}$ are presented in fig.\ \ref{fig_AmplParam}. On the upper pionic mode we instead parametrize the amplitudes as \begin{eqnarray} \sum X^{\tilde{\pi}_2}_{\Delta N^{-1}}(\mathbf{q};\rho) & \approx & C_N'[X^{\tilde{\pi}_2}_{ \Delta N^{-1}}] \sqrt{ f_2(q,\, \vec{C}'[X^{\tilde{\pi}_2}_{\Delta N^{-1}}]) } \label{eq_XDh-upp-parm} \\ \sum X^{\tilde{\pi}_2}_{ N N^{-1}}(\mathbf{q};\rho) & \approx & C_N'[X^{\tilde{\pi}_2}_{ N N^{-1}}] \sqrt{ 1- f_2(q,\, \vec{C}'[X^{\tilde{\pi}_2}_{ N N^{-1}}]) }\ , \label{eq_XNh-upp-parm} \end{eqnarray} where \begin{equation} f_2(q,\vec{C}') = \left[ 1 + \exp \left( C_0' + C_1' \, cq + C_{-1}' / cq \right) \right]^{-1}\ , \label{eq_f2(q,rho)-parm} \end{equation} with the density-dependent coefficients $\vec{C}'$, displayed in fig.\ \ref{fig_AmpuParam}. The amplitudes $Z$ of the pion component are obtained from the parametrization of the squared amplitudes, eqs.\ (\ref{eq_Zl2-parm}) and (\ref{eq_Zu2-parm}) below, as $Z = \sqrt{Z^2}$. The total $\Delta$ decay width, has apart from the partial contributions $\tilde{\Gamma}_\Delta^j$, also the partial contributions $\Gamma_\Delta^{N N^{-1}}$ and $\Gamma_\Delta^{\Delta N^{-1}}$. The partial width $\Gamma_\Delta^{N N^{-1}}$ gives the probability for the $\Delta$ to decay into a nucleon and a $N N^{-1}$ state. In a transport description, this implies that we initially have a $\Delta$ and after the decay process we have two nucleons above the Fermi surface and a hole left in the Fermi sea. But this is the same process as if the $\Delta$ would collide with a nucleon below the Fermi surface to give two nucleons above the Fermi surface. This process is normally already included in the collision term in a standard transport description, and the probability for such a collision is given by the cross section for the process $\Delta + N \rightarrow N + N$. In a transport description it is therefore not correct to both include a $\Delta$ decay according to $\Gamma_\Delta^{N N^{-1}}$ and a collision term with $\Delta + N \rightarrow N + N$. Instead, the correct procedure should be to exclude $\Gamma_\Delta^{N N^{-1}}$ and modify the cross section $\sigma(\Delta + N \rightarrow N + N)$ to be the in-medium cross section. Calculations of such in-medium cross sections was discussed in ref.\ \cite{main}. In the same way, $\Gamma_\Delta^{\Delta N^{-1}}$ should be excluded in a transport description, and $\sigma(\Delta + N \rightarrow \Delta + N)$ be the in-medium cross section. Although the collective pionic modes can thus be effectively treated as ordinary particles, the fact that their wave functions contain components from $\pi$, $NN^{-1}$ and $\Delta N^{-1}$ states makes it difficult to picture them in a physically simple manner. Fortunately, their specific structure is not important for the transport process, as as long as these quasiparticles remain well inside the nuclear medium. First when such a quasiparticle penetrates a nuclear surface and emerges as a free particle is it physically meaningful to determine what kind of real particle it is. The gradual transformation of the collective quasiparticle occurs automatically within the formalism, because as the density is lowered, $\rho \rightarrow 0$, a pionic mode will acquire 100\% of either the pion component or the $\Delta N^{-1}$ component, depending on $\omega$ and $q$. That is to say, it will turn into either a free pion or an unperturbed $\Delta N^{-1}$ state. The squared amplitudes have their values between zero and unity and we therefore employ the parametrizations \begin{eqnarray} Z^{\tilde{\pi}_1}(\mathbf{q};\rho)^2 & \approx & f_2(q,\, \vec{C}''[Z^{\tilde{\pi}_1}]), \label{eq_Zl2-parm} \\ Z^{\tilde{\pi}_2}(\mathbf{q};\rho)^2 & \approx & 1-f_2(q,\, \vec{C}''[Z^{\tilde{\pi}_2}]) \label{eq_Zu2-parm} \\ \sum X^{\tilde{\pi}_1}_{\Delta N^{-1}}(\mathbf{q};\rho)^2 & \approx & 1-f_2(q,\, \vec{C}''[X^{\tilde{\pi}_1}_{\Delta N^{-1}}]) \label{eq_XDhl2-parm} \\ \sum X^{\tilde{\pi}_2}_{\Delta N^{-1}}(\mathbf{q};\rho)^2 & \approx & f_2(q,\, \vec{C}''[X^{\tilde{\pi}_2}_{\Delta N^{-1}}])\ , \label{eq_XDhu2-parm} \end{eqnarray} with $f_2(q,\vec{C}'')$ from eq.\ (\ref{eq_f2(q,rho)-parm}), while the sum of the squared $N N^{-1}$ amplitudes are obtained from the normalization, \ie\ all squared amplitudes sum up to unity. The density-dependent coefficients $\vec{C}''$, are presented in fig.\ \ref{fig_Amp2Param}. There remains the practical problem of how to represent an unperturbed $\Delta N^{-1}$ state when $\rho \rightarrow0$. However, as will be discussed in section \ref{sec_Qpions}, only a very small fraction of the pionic modes (vanishing for a stationary density profile) will emerge as unperturbed $\Delta N^{-1}$ states. Note that this approach is different from earlier works \cite{Giessen,Texas,KochPriv} where the nature of the pionic mode was determined already in the creation process, \ie\ when the mode was created it was determined whether it represented a free pion or an unperturbed $\Delta N^{-1}$ state. This difference in approach will have crucial effects for the collective modes that escape the system, as will be seen in the next sections. \section{ Quasipions at the nuclear surface } \label{sec_Qpions} In previous works the quasipions at a surface have been treated in various approximations. In ref.\ \cite{Giessen} effective dispersion relations were introduced, corresponding to modes with either 100\% pionic or $\Delta N^{-1}$ component. The pionic mode was then propagated as a quasipion, emerging as a free pion at the surface, while the $\Delta N^{-1}$ mode was used to derive a $\Delta$ potential for the uncoupled $\Delta$s. In this way only pions and $\Delta$s escape the system, but the drawback is that the effective dispersion relations are quite distorted compared to the original ones, and that the in-medium effects seem to be over-estimated by allowing all uncoupled $\Delta$s propagate in a collective potential. In \cite{Texas} both collective modes were propagated, and hence some modes escape the system as unperturbed $\Delta N^{-1}$ states. This was effectively taken care of by converting these modes to free $\Delta$s (neglecting baryon number conservation), with the justification that the number of modes escaping the system as unperturbed $\Delta N^{-1}$ states were found to be small. We will in this paper report on alternative ways to treat the modes at the surface based on ref.\ \cite{main}, and how this treatment can improve the descriptions in refs.\ \cite{Giessen,Texas}. We first consider the simplified case when the nuclear surface is stationary \ie\ $\rho(\mathbf{r},t) = \rho(\mathbf{r},0)$ for all times $t$. This means that there is no explicit time dependence in the effective Hamiltonians in equation (\ref{eq_Hqpion}), and thus the energies of the collective pionic modes are conserved. For a pionic mode with energy $\hbar \omega_\nu(\mathbf{q};\rho)$ propagating in a varying density this means that the momentum $\mathbf{q}$ will change as $\rho$ changes. That is to say, the pionic mode effectively feels a potential. In addition to momentum changes, also the amplitudes $X^\nu_{jk}(\mathbf{q};\rho)$, $Z^\nu(\mathbf{q};\rho)$ and $W^\nu(\mathbf{q};\rho)$ of the baryon-hole and pion components will change as the density changes. In the limit of vanishing density either $X^\nu_{\Delta N^{-1}}$ or $Z^\nu$ will turn to unity, depending on the mode $\nu$ and its energy and momentum. In the latter case no problems arise, the collective mode has simply been converted to a free pion escaping the system. However, in the former case there is some inconsistency in the formalism since there are no hole states in vacuum ($\rho=0$). In a quantal description of spin-isospin modes propagating at a surface different scenarios could emerge. There is some small probability that the mode could be reflected at the surface. Alternatively the mode could break up in an uncoupled $\Delta$ escaping the system, with the hole is trapped inside the nucleus. In a transport description this could be handled by allowing the test particle representing the mode $\nu$ to absorb a nearby nucleon, converting it into a $\Delta$ isobar. We consider in this section the quasipions at a surface propagating from normal nuclear density to vacuum. In this case the amplitude $W^\nu$ will be very small and we will therefore omit it in the qualitative discussion of this section. In fig.\ \ref{fig_w12-rho}$a$ we present the energy $\hbar \tilde{\omega}_1(\mathbf{q})$ of the lower collective mode $\tilde{\pi}_1$ for different densities in the range $0.1 \rho^0 \leq \rho \leq \rho^0$. The line closest to the free pion relation (dotted line) represents the dispersion relation at the lowest density. As the density is increased the energy relation is (for each fixed $q$) lowered. For small $q$ values the mode is completely dominated by the pion component. As $q$ is increased also the $\Delta N^{-1}$ components starts to contribute substantially (although to different extent depending on the density). In the approximate range $300\ \MeV/c \leq q \leq 400\ \MeV/c$ (depending on density) the mode enters the $\Delta N^{-1}$ continuum and changes character from collective mode to non-collective. We will in this section discuss four different examples, suitably chosen to illustrate the main features. \subsection{ Lower pionic mode } \label{sec_QpionsLow} In our first example we consider the mode $\tilde{\pi}_1$ created at normal density with energy $\hbar \tilde{\omega}_1 = 200\ \MeV$ propagating towards vacuum without any interactions. Initially this mode has the following characteristics: \[ \begin{array}{cccc} q \approx 220\ \MeV/c & |Z|^2 \approx 0.72 & \sum_{\Delta N^{-1}} |X_{\Delta N^{-1}}|^2 \approx 0.25 & \sum_{N N^{-1}} |X_{N N^{-1}}|^2 \approx 0.03\ . \end{array} \] As the density decreases it will, due to the energy conservation, follow the path indicated by the dashed line in fig.\ \ref{fig_w12-rho}$a$, from $q \approx 220\ \MeV/c$ to $q \approx 150\ \MeV/c$. In fig.\ \ref{fig_Amp-w}$a$ we see how the squared amplitudes vary with the density for this particular energy of the mode $\tilde{\pi}_1$. As seen in fig.\ \ref{fig_Amp-w}$a$ only the pion component remains, as the zero density limit is reached. Thus in this particular case, the mode will escape the system as a free pion. In fig.\ \ref{fig_DhAmp-w}$a$ we show the squared amplitudes of the individual $\Delta N^{-1}$ amplitudes for the energy $\hbar \tilde{\omega}_1 = 200\ \MeV$ of the mode $\tilde{\pi}_1$. We see that the mode is collective at all densities. In the discussion so far we have not addressed the creation process. How probable is it that we produce the mode $\tilde{\pi}_1$ at the particular energy $\hbar \tilde{\omega}_1 = 200\ \MeV$? This information is given by the partial $\Delta$ decay width to the mode $\tilde{\pi}_1$. In fig.\ \ref{fig_Gam_w-rho}$a$ we display the partial $\Delta$ decay width, $\tilde{\Gamma}_\Delta^1$, for different densities, as a function of the energy of the emitted quasipion $\tilde{\pi}_1$. Note that the width displayed in fig.\ \ref{fig_Gam_w-rho}$a$ is for a $\Delta$ at rest, and that the Pauli blocking of the emitted nucleon has not been taken into account (the Pauli blocking is treated explicitly in a transport description). As seen in fig.\ \ref{fig_Gam_w-rho}$a$ the width is quite substantial at the quasipion energy $\hbar \tilde{\omega}_1 = 200\ \MeV$, about $80\ \MeV$ compared to the free width, which is about $30\ \MeV$ at this energy. Thus it is quite probable for a $\Delta$ to decay to the mode $\tilde{\pi}_1$ around this quasipion energy. Apart from penetrating the surface there is also some probability for the mode $\tilde{\pi}_1$ to be reflected at the surface. This is difficult to exactly predict since the reflection coefficient, ${\cal R}$, will depend on the actual density profile. However a first estimate can be obtained by considering a one-dimensional scenario with a density profile that corresponds to a Wood-Saxon potential with a surface thickness $a=0.65$ fm. The height of the potential is then given by the change in the momentum of the pionic mode, and for the case of $\hbar \tilde{\omega}_1 = 200\ \MeV$ we obtain ${\cal R} \approx 0.017$, see further the discussion in appendix \ref{sec_Refl}. This number is very small, which implies that the reflection can practically be neglected for this particular case. However, in other situations where the momentum change is different or the density profile is sharper, the reflection coefficient may be larger. We have therefore devoted appendix \ref{sec_Refl} to discuss how effective local reflection coefficients can be obtained and implemented in transport descriptions. In our second example we also consider the mode $\tilde{\pi}_1$, but now with energy $\hbar \tilde{\omega}_1 = 295\ \MeV$. Initially this mode will now have the characteristics: \[ \begin{array}{cccc} q \approx 450\ \MeV/c & |Z|^2 \approx 0.0 & \sum_{\Delta N^{-1}} |X_{\Delta N^{-1}}|^2 \approx 0.94 & \sum_{N N^{-1}} |X_{N N^{-1}}|^2 \approx 0.06\ . \end{array} \] As the density decreases it follows the path indicated by the dot-dashed line in fig.\ \ref{fig_w12-rho}$a$, from $q \approx 450\ \MeV/c$ to $q \approx 250\ \MeV/c$. In fig.\ \ref{fig_Amp-w}$b$ we see how the squared amplitudes vary with the density for this particular energy of the mode $\tilde{\pi}_1$. As seen in fig.\ \ref{fig_Amp-w}$b$ also in this case only the pion component remains, as the zero density limit is reached, although $\tilde{\pi}_1$ initially was completely dominated by the $\Delta N^{-1}$ component. Thus also in this case the mode will escape the system as a free pion. In fig.\ \ref{fig_DhAmp-w}$b$ we show the squared amplitudes of the individual $\Delta N^{-1}$ amplitudes for the energy $\hbar \tilde{\omega}_1 = 295\ \MeV$ of the mode $\tilde{\pi}_1$. Here the situation is different from the case at $\hbar \tilde{\omega}_1 = 200\ \MeV$, since initially the mode is dominated by a single $\Delta N^{-1}$ component, \ie\ the mode is non-collective. However as the density is lowered the strength is spread over more $\Delta N^{-1}$ components and at low densities the mode is completely collective. {}From fig.\ \ref{fig_Gam_w-rho}$a$ we see that the partial width $\tilde{\Gamma}_\Delta^1$ is very close to zero at the quasipion energy $\hbar \tilde{\omega}_1 = 295\ \MeV$ (dot-dashed curve) at normal nuclear density. Thus a $\Delta$ at normal nuclear density will not decay to a quasipion with this energy. The partial $\Delta$ width becomes very small because the collective strength, as well as the pion component, is negligible in this case. Although the mode at this energy cannot be created at normal nuclear density, it may still be created at lower densities, as can be seen in fig.\ \ref{fig_Gam_w-rho}$a$. Making the same assumptions as in the first example, we find that the reflection coefficient, becomes smaller than $10^{-4}$ for this case. We thus conclude that the lower pionic mode penetrating the surface will always emerge as a free pion. For low energies this is natural because as the density decreases the dispersion relation of the pionic mode approaches the free pion relation. The cases of sufficiently large energy to approach the unperturbed $\Delta N^{-1}$ branch as the density is lowered will never occur since no such modes will be created from a decaying $\Delta$, because the partial $\Delta$ width will be zero. \subsection{ Upper pionic mode } \label{sec_QpionsUpp} Also on the upper collective mode, $\tilde{\pi}_2$ a similar effect will occur. However some properties are somewhat different so we will therefore illustrate also this case with two typical examples. In our third example we thus consider the mode $\tilde{\pi}_2$, with energy $\hbar \tilde{\omega}_2 = 320\ \MeV$. Note that at this energy the mode can only exist at densities up to about $0.5 \rho^0$. Initially, at $\rho = 0.5 \rho^0$, this mode will have the characteristics: \[ \begin{array}{cccc} q \approx 0\ \MeV/c & |Z|^2 \approx 0.0 & \sum_{\Delta N^{-1}} |X_{\Delta N^{-1}}|^2 \approx 1.0 & \sum_{N N^{-1}} |X_{N N^{-1}}|^2 \approx 0.0\ . \end{array} \] As the density decreases we follow the path indicated by the dashed line in fig.\ \ref{fig_w12-rho}$b$, from $q \approx 0\ \MeV/c$ to $q \approx 270\ \MeV/c$. Note that on the upper collective mode the momentum increases as the density decreases, corresponding to a negative potential step. In fig.\ \ref{fig_Amp-w}$c$ we see how the squared amplitudes vary with the density for this particular energy of the mode $\tilde{\pi}_2$. As seen in fig.\ \ref{fig_Amp-w}$c$, contrary to previous examples, only the $\Delta N^{-1}$ component remains, as the zero density limit is reached. In fig.\ \ref{fig_DhAmp-w}$c$ we show the squared amplitudes of the individual $\Delta N^{-1}$ amplitudes for the energy $\hbar \tilde{\omega}_2 = 320\ \MeV$ of the mode $\tilde{\pi}_2$. Initially the mode is collective, but as the density is lowered the mode becomes more and more non-collective. {}From fig.\ \ref{fig_Gam_w-rho}$b$ we however see that the partial width $\tilde{\Gamma}_\Delta^2$ actually is zero at the quasipion energy $\hbar \tilde{\omega}_2 = 320\ \MeV$ (dot-dashed curve) at all densities. Thus a $\Delta$ will not decay to a quasipion with this energy. A mode at this energy with a very low momentum at half nuclear density, which could occur in a time-dependent density, could have a very large reflection coefficient, approaching unity as the initial quasipion momentum approaches zero. In our fourth and last example in this section we consider the mode $\tilde{\pi}_2$ with energy $\hbar \tilde{\omega}_2 = 380\ \MeV$. Initially this mode will now have the properties: \[ \begin{array}{cccc} q \approx 170\ \MeV/c & |Z|^2 \approx 0.12 & \sum_{\Delta N^{-1}} |X_{\Delta N^{-1}}|^2 \approx 0.88 & \sum_{N N^{-1}} |X_{N N^{-1}}|^2 \approx 0.0\ . \end{array} \] As the density decreases we follow the path indicated by the dot-dashed line in fig.\ \ref{fig_w12-rho}$b$, from $q \approx 170\ \MeV/c$ to $q \approx 350\ \MeV/c$. In fig.\ \ref{fig_Amp-w}$d$ we see how the squared amplitudes vary with the density for this particular energy of the mode $\tilde{\pi}_2$. As seen in fig.\ \ref{fig_Amp-w}$d$ in this case only the pion component remains, as the zero density limit is reached, although $\tilde{\pi}_2$ initially was dominated by the $\Delta N^{-1}$ component. Thus also in this case the mode will escape the system as a free pion. In fig.\ \ref{fig_DhAmp-w}$d$ we show the squared amplitudes of the individual $\Delta N^{-1}$ amplitudes for the energy $\hbar \tilde{\omega}_2 = 380\ \MeV$ of the mode $\tilde{\pi}_2$. We see that the mode at all densities is completely collective. {}From fig.\ \ref{fig_Gam_w-rho}$b$ we see that the partial width $\tilde{\Gamma}_\Delta^2$ is quite substantial at the quasipion energy $\hbar \tilde{\omega}_2 = 380\ \MeV$ (dot-dashed curve) also at normal nuclear density. The reflection coefficient, making the same assumptions as in the first example, becomes for this case smaller than $10^{-3}$. We thus conclude that also the upper pionic mode penetrating the surface will always emerge as a free pion. For high energies this is natural because as the density decreases the dispersion relation of the pionic mode approaches the free pion relation. The cases of low energy when the unperturbed $\Delta N^{-1}$ branch is approached as the density is lowered, will never occur since no such modes will be created from a decaying $\Delta$, because the partial $\Delta$ width will be zero. \subsection{ Refined scenarios } \label{sec_QpionsOther} If the surface changes with time the energy of the pionic mode need not to be conserved, and there is some small possibility for the pionic mode to end up as an unperturbed $\Delta$-hole state. The actual fraction of such modes is hard to estimate without an explicit transport simulation, but based on the scenario for the time-independent density profile, it is reasonable to expect that only a very small fraction of the pionic modes will end up as unperturbed $\Delta$-hole states in vacuum. In a quantum description such a mode could be either reflected at the surface, or the mode could break up into an uncoupled $\Delta$ and hole, where the hole is trapped inside the nucleus, and the $\Delta$ escape the system as a free $\Delta$. Based on the results discussed in the explicit examples of this section, and the presentation in appendix \ref{sec_Refl} we expect that the reflection at the surface will be very small. In a transport description the reflection can be incorporated by a local transmission coefficient as discussed in appendix \ref{sec_Refl}. If the pionic mode is not reflected at the surface, and its amplitude approaches 100\% of the $\Delta N^{-1}$ component as the density approaches zero, the mode should thus break up into an uncoupled $\Delta$ and hole. In a transport simulation this could be practically handled by allowing the pionic mode to absorb a nearby nucleon, forming an uncoupled $\Delta$, when the density falls below a specified value. This prescription has some quantum mechanical justification by the fact that at very low density a wave packet representing the pionic mode, will not be very well localized, but instead have a large spatial spread. \section{Summary} \label{sec_Sum} In-medium properties obtained in an infinite stationary system consisting of interacting nucleons, nucleon resonances and mesons, can be incorporated into transport descriptions by a local density approximation. While such a prescription is rather straightforward to implement in the interior regions of the nuclear system, conceptual problems exist at the nuclear surface. When the nuclear density approaches zero, collective mesonic modes formed in the medium have to be converted to real particles in vacuum. The problems arise since some collective modes (e.g. $\Delta N^{-1}$-like) may exists in the infinite stationary system at arbitrary low (but non-vanishing) density, but no corresponding real particle exists in vacuum. This problem has been apparent in previous works \cite{Giessen,Texas} where collective modes have been incorporated into transport descriptions as quasimesons. The character of the quasimesons (\ie\ realization in vacuum) were in those works determined already at the time of creation. Based on the formalism of ref.\ \cite{main}, we have in this paper employed a more elaborate $\pi + N N^{-1} + \Delta N^{-1}$ model (relative to the works \cite{Giessen,Texas}) to investigate a somewhat different treatment of the collective pionic modes at a nuclear surface. In this formalism we have obtained not only density dependent dispersion relations of the pionic modes, but also density dependent amplitudes of the components constituting the pionic mode. These quantities are conveniently parametrized with density dependent parameters, in section \ref{sec_DispAmp}. For the transport process it is not needed to determine the character of the pionic modes until they penetrate the surface and emerge as free particles. This is automatically determined within our formalism from the amplitudes at zero density. We have further showed in section \ref{sec_Qpions} that for a stationary density profile, the conservation of the energy of the pionic mode and the partial $\Delta$ decay width, together leads to the fact that only real pions are realized as free particles when the pionic mode penetrates the surface. Note that this finding is different from earlier works, and it demonstrates the importance of deriving dispersion relations and partial $\Delta$ widths consistently within a realistic\footnote{By ``realistic'' we here mean that the $\Delta N^{-1}$ model contains a continuum of $\Delta N^{-1}$ and $N N^{-1}$ states, as compared to the more simple two-level $\Delta N^{-1}$ model used for example in refs.\ \cite{Giessen,Texas}. } $\Delta N^{-1}$ model. In a more refined scenario, where the density changes with time, deviations from this picture can be expected and also the unperturbed $\Delta N^{-1}$ component may be realized in the limit of vanishing density. The actual fraction of such modes is hard to estimate without an explicit transport simulation, but based on the arguments in section \ref{sec_Qpions} for the stationary surface, we expect this fraction to be very small. For the rare cases when the unperturbed $\Delta N^{-1}$ component is realized in the limit of vanishing density, the pionic mode must be converted to real particles. In an extended description this could be made by allowing the mode to break up into an uncoupled $\Delta$ and a $N^{-1}$, where the hole remains trapped in the nuclear system and the $\Delta$ escapes the system. Based on our formalism, this seems to be the most probable scenario, since the collective strength disappears on these modes as the density approaches zero. This could be implemented in a transport simulation by letting the pionic mode absorb a nearby nucleon to form an uncoupled $\Delta$, as may be justified by the quantum-mechanical feature that the wave packet representing the pionic mode is not well localized at very low density. Alternatively the pionic modes could in a quantal description be reflected at the surface. We have investigated the reflection and transmission probabilities for the collective modes in a simplified one-dimensional scenario, where the modes propagate perpendicular to the surface. Exploring typical scenarios, we have found that (with only very few exceptions) the reflection of the pionic modes will be smaller than a few percent, however, in appendix \ref{sec_Refl} we have suggested how the reflection and transmission probabilities could be incorporated into the transport descriptions by using approximative local transmission coefficients. \\ \noindent Stimulating discussions with Volker Koch are acknowledged. This work was supported by the Swedish Natural Science Research Council, and by the Director, Office of Energy Research, Office of High Energy and Nuclear Physics, Nuclear Physics Division of the U.S. Department of Energy under Contract No.\ DE-AC03-76SF00098.
\section{Introduction} The problem of the gluon vacuum in QCD has a long history \cite{savvidy,coleman,leutwyler,belavin,zakharov,shuryak}. It is well known that at low energies perturbation theory does not apply since the coupling constant becomes large and the perturbative vacuum becomes unstable due to gluon self-interactions \cite{savvidy}. These self-interactions can lead to a 'reconstruction' of the vacuum and to the appearance of a condensate \cite{nn,umt}. The presence of the condensate is important for the physical properties of the low energy sector of QCD. Of particular interest is the modification of the quasiparticle spectrum and the occurence of massive collective excitations to be considered in this work. In the literature two different kinds of such condensates have been considered, the {\it coherent} and the {\it squeezed} one. In the coherent condensate gluon field excitations are found by a {\it transitive} transformation, i.e. shifting the fields to the solution of the classical equations \cite{savvidy,coleman,leutwyler} (e.g. instantons \cite{belavin,zakharov,shuryak}). It is characterized by the condensation of single gluons and thus by a nonvanishing vacuum expectation value of the gluon field $A$: \begin{equation} <A>\ \ \neq\ \ 0~. \end{equation} In the squeezed condensate on the other hand the gluon states are constructed by a {\it multiplicative} Bogoliubov transformation of the gluon fields \cite{celenza,biro1,biro2,mishra1,mishra2}. The squeezed vacuum is characterized by the condensation of colourless, scalar gluon pairs and could thus be realized through \begin{equation} <A>\ \ =\ \ 0\ ,\ \ <A^2>\ \ \neq 0\ . \end{equation} There has been a lot of activity to construct a stable coherent vacuum in the gluon sector of QCD \cite{savvidy,coleman,leutwyler}. The problem in this case is that there are no stable quasiclassical solutions to the Yang-Mills equations in Minkowski space \cite{coleman,leutwyler}. In recent years the squeezed condensate (called here {\it Bogoliubov condensate}) has become a topic of great interest, see e.g. \cite{celenza,biro1,biro2,mishra1,mishra2}. Its investigation for non-Abelian fields faces the following problems: \begin{enumerate} \item One has to find the adequate degrees of freedom to construct the gluon condensation in a squeezed vacuum. \item Since the squeezed vacuum is most naturally described in the Hamiltonian formalism one has to extract the gauge invariant oscillator - like field variables from the non-Abelian QCD action. Recall that by using different gauges for instance Biro \cite{biro1,biro2} and Mishra \cite{mishra1,mishra2} get different results for the condensate. \item The condensate leads to spontaneous gauge symmetry breaking with the appearance of massive gluons. This has to be accompanied by the corresponding generation of the longitudinal components for the massive gluon fields. Recent papers on ''squeezed'' gluon states \cite{biro1,biro2,mishra1,mishra2} indeed obtain a constituent gluon mass. However, since they fix the gauge, their gluon fields do not have a longitudinal component such that they are unable to provide the number of degrees of freedom required for the description of a massive gluon vector field \cite{slavnov}. \end{enumerate} With respect to the last topic, we mention that mass generation for gauge fields via spontaneous gauge symmetry breaking is a general problem familiar also from other gauge field theories. In the unified theory of electroweak interactions a consistent description of massive vector fields can be given due to the presence of a scalar Higgs field in the Lagrangian which generates the longitudinal component of the massive gauge bosons W and Z. The very interesting alternative, that the gauge bosons obtain their mass by spontaneous gauge symmetry due to radiative corrections, without introduction of an external Higgs field, has been proposed by Coleman and Weinberg \cite{cw}. This possibility is of great importance since for SU(N) gauge theories the introduction of an external Higgs field does not lead to a mass term for the gluons, as was shown by Georgi and Glashow \cite{Georgi}. The concept of spontaneous gauge symmetry breaking by radiative corrections is therefore very attractive for the case of QCD and has been followed up to now \cite{biro1,biro2,mishra1,mishra2}. In the present work we study the possibility of gluon condensation in a squeezed vacuum in view of the Bogoliubov model \cite{nn} of the weakly nonideal Bose gas. In particular we investigate the influence of a squeezed condensate on the gluon quasiparticle spectrum in the low-energy region of QCD. Note that the Bogoliubov theory was the first to explain the experimentally observable spectrum of collective excitations of superfluid $^4$He. This spectrum cannot be obtained by resummations of the conventional perturbation theory series. As Bogoliubov has shown, the collective exitations are determined by the ''condensate'' of particles with zero momentum and finite density. A first connection between Bogoliubov condensation and the squeezed vacuum in field theory (massless $\lambda \phi^4$) was made by Castorina and Consoli \cite{CC}, see also \cite{PaBla}. An application of the concept of the Bogoliubov model to QCD, however, has to our knowledge not been carried out by now. In a first attempt to generalize the Bogoliubov model to QCD we use the infrared singularity of massless theories to squeeze the zero momentum mode which leads to massive gluonic quasiparticles in the nonzero momentum sector. The free squeezing parameter is fixed from the $\eta-\eta'$ mass difference. Our coresponding value for the gluon condensate is in reasonable agreement with that obtained by Shifman, Vainshtein and Zakharov \cite{zakharov} which supports our semi-phenomenological approach. Concerning the above mentioned problems (2) and (3), we try to solve the problem of the appearance of a constituent gluon mass using a gauge invariant scheme for the elimination of the unphysical components of the gluon vector field \cite{dirac,perv1,perv2,fj,kp} which does not require the gauge-fixing as initial supposition. This scheme is based on the construction of projection operators by formally solving the Gauss law constraint. We show that these projectors are destroyed by the interaction of gluons with the squeezed vacuum. As result a constituent gluon mass appears together with the necessary longitudinal components. This is the central result of our paper and is quite in analogy to spontaneous chiral symmetry breaking in the quark sector \cite{ad,lopr,pk2s}. There, the appearance of constituent quark masses due to the interaction of quarks with the squeezed vacuum is accompanied by the destruction of the chiral projection operator which leads to the necessary increase of the number of spinor field components from two (Weyl spinors) to four (Dirac spinors). The paper is organized as follows: In Section \ref{sec:field}, the Bogoliubov model of a weakly nonideal Bose gas is generalized to field theory. We give a field theoretical description of the condensation phenomenon by the use of the squeezed vacuum and discuss the conventional local $\lambda \phi^4$ theory. The relation to the Bogoliubov model for the weakly nonideal Bose gas is given in Appendix \ref{app:bogmod}. In Section \ref{sec:qcd} the homogeneous colourless Bogoliubov condensate of gluons is introduced in QCD where the unphysical degrees of freedom are eliminated by applying projection operators instead of fixing a gauge. We also discuss spontaneous gauge symmetry breaking and the corresponding occurrence of a massive gluon quasiparticle spectrum. In Section \ref{sec:app} the squeezing parameter is fixed >from the $\eta' - \eta$ mass difference. In Section \ref{sec:end} we present the conclusions. \section{Bogoliubov condensation in quantum field theory} \label{sec:field} In order to introduce some notations and methods needed for our investigation of the squeezed condensate in the rather complicated QCD, we first consider massless $\lambda\phi^4$ theory with the Hamiltonian \begin{equation} H=\int d^3x[\pi(x)^2+\left(\partial_i\varphi(x)\right)^2+{\lambda\over 4!} \varphi^4(x)] \end{equation} as the simplest example of an interacting bosonic theory which is renormalizable. The theory is quantized by turning the classical fields $\varphi({\bf x},t),\pi({\bf x},t)$ to Schr\"odinger operators $\varphi({\bf x}),\pi({\bf x})$ and imposing the canonical commutation relations \begin{equation} [\pi({\bf x}),\varphi({\bf x}')]=-i\delta({\bf x}-{\bf x}')~. \end{equation} In the momentum representation defined by \begin{equation} \label{FouTra} \varphi_p={1\over\sqrt{V}}\int d^3x \mbox{\large e}^{i{\bf p}{\bf x}}\varphi({\bf x}),\ \ \ \ \pi_p={1\over\sqrt{V}}\int d^3x \mbox{\large e}^{i{\bf p}{\bf x}}\pi({\bf x})~, \end{equation} the Hamilton operator is \begin{equation} \label{hf1} :{ H}[\varphi,\pi]:= \frac{1}{2} \sum_p \left[:\pi_p \pi_{-p}: + p^2 :\varphi_p \varphi_{-p}: \right] + \frac{\lambda}{4!V} \sum_{p_1 p_2 p_3 p_4} \delta_{p_1+p_2+p_3+p_4,0} :\varphi_{p_1}\varphi_{p_2}\varphi_{p_3}\varphi_{p_4}: , \end{equation} with the commutation relations \begin{equation} \label{crphipi} [\pi_p,\varphi_{p'}]= - i\delta_{p,-p'} ~,~~ [\varphi_p,\varphi_{p'}]= [\pi_p,\pi_{p'}]= 0~. \end{equation} In (\ref{hf1}) we have introduced normal ordering with respect to the creation and annihilation operators $a_p, a_p^+$ defined according to \begin{eqnarray} \label{phipi} \varphi_p =\sqrt{\frac{1}{2 \tilde{\omega}(p)}} (a_p + a^+_{-p})\ , \;\ \ \ \pi_p = i \sqrt{\frac{\tilde{\omega}(p)}{2 }} (-a_p + a^+_{-p}) ~, \end{eqnarray} with an arbitrary function $\tilde{\omega}(p)$. The operators $a_p, a_p^+$ satisfy the commutation relations \begin{equation} \label{craa} [a_p, a_{p'}^+]= \delta_{p, p'}~,~~[a_p, a_{p'}]= [a_p^+, a_{p'}^+]= 0 ~. \end{equation} The corresponding vacuum $|0> $ is defined by $a_p |0> = 0$, and the Fock space is given as \begin{eqnarray} \label{fockstate} \{\,|\,\Phi >\} = \,|\,0 >;\;\;\; a^+_p |0 > = |\,p >, ... ~. \end{eqnarray} We note that the special choice $\tilde{\omega}(p)=|p|$ diagonalizes the free ($\lambda$ independent) part of the Hamiltonian. For this case, however, $a_0$ and $a_0^+$ are not defined which corresponds to the well-known infrared singularity of massless theories. Generalizing the Bogoliubov model to field theory we should use the infrared singularity of massless theories to squeeze the zero mode by populating it macroscopically with massless particles. Then, we diagonalize the nonzero mode single particle part of the resulting squeezed Hamiltonian by changing from particles to quasiparticles whose dispersion relation $\tilde{\omega}(p)$ is finally determined selfconsistently. For the simple case of $\lambda\phi^4$ this has been carried out in detail \cite{PaBla}. Similar to the Bogoliubov model this leads to renormalization of the bare parameters like the coupling constant $\lambda$. For the time being we leave $\tilde{\omega}(p)$ open and suppose that the vacuum of the theory (\ref{hf1}) contains a large number of quasiparticles with zero momentum ($p=0$). We construct this vacuum using the unitary squeezing operator \begin{equation} \label{ub} U_B (\varphi_0,\pi_0) = \exp \left (i\frac{f_0}{2} (\pi_0 \varphi_{0} + \varphi_0 \pi_{0}) \right )~, \end{equation} where $f_0$ is a very large parameter to be fixed later. The operator $U_B$ transforms the Fock space of states to the Bogoliubov space of states \begin{eqnarray} \label{bogstate} |\, \Phi_B >\equiv U_B^{-1} \,|\, \Phi > ~. \end{eqnarray} In quantum optics these states are called 'squeezed states', see e.g. \cite{optics}. Applying the unitary transformation (\ref{ub}), we can define the new field operator $\varphi^B_0$ and its momentum $\pi^B_0$ by means of \begin{eqnarray} \label{bt} \varphi^B_0 = U_B ^{-1}\varphi_0 U_B = {\rm e}^{-f_0} \varphi_0~,\nonumber\\ \\ \pi^B_0 = U_B ^{-1}\pi_0 U_B = {\rm e}^{f_0} \pi_0 \nonumber~, \end{eqnarray} which satisfy the same algebra of commutation relations as the initial ones (\ref{crphipi}). We shall now carry out the squeezing of the zero mode part of the Hamiltonian by applying a Wick reordering procedure to the Bogoliubov vacuum $|0_B>$. Note that under the squeezing transformation (\ref{bt}) the contraction of a pair of field operators is left invariant, \begin{equation} C= <0 \,|\,\varphi_{0}\varphi_{0} \,|\,0> = <0_B\,|\,\varphi^B_{0} \varphi^B_{0}\,|\,0_B>~. \end{equation} The normal ordering of the Bogoliubov fields $\varphi_B$ with respect to the Bogoliubov vacuum (which is denoted as $::\varphi^B_{0} \varphi^B_{0} ::$) has the same form as the normal ordering of the original fields with respect to the Fock vacuum (\ref{fockstate}) \begin{eqnarray} \label{no} C = \varphi_{0} \varphi_{0} - :\varphi_{0} \varphi_{0}: = \varphi^B_{0} \varphi^B_{0} - :: \varphi^B_{0} \varphi^B_{0} ::~~. \end{eqnarray} To reorder the Hamiltonian (\ref{hf1}) with respect to the new vacuum $|0_B >$ we use Eqs.(\ref{bt}) and (\ref{no}). Reordering of the quadratic term gives \begin{equation} \label{bc} :\varphi_{0} \varphi_{0}: = ::\varphi_{0}^B \varphi_{0}^B::~ \mbox{\large{e}}^{2f_{0}}~ + ~\tilde{C} ~, \end{equation} where \begin{eqnarray} \label{bfi} \tilde{C}= C \left (\mbox{\large{e}}^{2f_{0}} - 1 \right )~. \end{eqnarray} Analogously we have \begin{equation} :\pi_{0} \pi_{0}: = ::\pi_{0}^B \pi_{0}^B::~ \mbox{\large{e}}^{-2f_{0}}~+~ C^{\pi} \left (\mbox{\large{e}}^{-2f_{0}} - 1 \right ), \end{equation} with $ C^{\pi}=$ Reordering of the quartic term gives \begin{eqnarray} :\varphi_{0}\varphi_{0}\varphi_{0}\varphi_{0}: &=& ::\varphi_{0}^B \varphi_{0}^B \varphi_{0}^B \varphi_{0}^B:: \mbox{\large{e}}^{4f_{0} } +:: \varphi_{0}^B \varphi_{0}^B:: \mbox{\large{e}}^{2f_{0}} \tilde{C} +(5\,\, {\rm permutations})\nonumber\\ & & +\tilde{C}^2 + (2 \,\,{\rm permutations})~. \end{eqnarray} For applications in QCD we quote here also the general Wick reordering result for any polynomial $F(\varphi_0)$ \begin{equation} :F(\varphi_0):=\exp\{{1\over 2}\tilde{C} {d^2\over d b^2}\}::F(\varphi^B_0\mbox{\large{e}} ^{f_0}+b)::\Big|_{b=0}~. \end{equation} As result of the reordering of $H=H_0+H'$, where $H'$ is the Hamiltonian of nonzero momentum excitations ($p\neq 0$), we obtain \begin{eqnarray} :H:&=& ::H_0(\phi_0):: + :H':~,\nonumber\\ \label{hfb} :H'[\varphi,\pi]: &=& E_0 + :H^{(2)}[\varphi,\pi]: + :H^{(4)}[\varphi]:~, \end{eqnarray} where \begin{eqnarray} \label{h0} E_0&=& 3{\lambda\over 4!V}\tilde{C}^2 + C^{\pi}\\ :{H}^{(2)}[\varphi,\pi]: &=& \frac{1}{2}\sum_{p\neq 0} \left\{:\pi_{p} \pi_{-p}: + \left[p^2 + {\lambda\over 2V} \tilde{C} _{00}\right] :\varphi_{p} \varphi_{-p}: \right\} ~,\\ :{ H}^{(4)}[\varphi]: &=&{\lambda\over 4!V} \sum_{p_1,p_2,p_3,p_4\neq 0}\delta_{p_1+p_2+p_3+p_4,0} :\varphi_{p_1}\varphi_{p_2}\varphi_{p_3}\varphi_{p_4} :~. \end{eqnarray} The zero momentum operator $::H_0(\phi_0)$ containing terms proportional to $::\varphi_{0}^B \varphi_{0}^B::$ and $::\varphi_{0}^B \varphi_{0}^B \varphi_{0}^B \varphi_{0}^B::$ describes excitations of the condensate and has not been written explicitely here. Note that for very large $f_0$ the second term in the expression (\ref{h0}) for $E_0$ is much smaller than the first one (cf. (\ref{bfi})) and can be neglected. Hence we find a condensate energy density \begin{equation} \label{Condene} \epsilon_0\equiv{E_0\over V}= {\lambda\over 8}\left({\tilde{C}\over V}\right)^2 \end{equation} and a bosonic quasiparticle mass $m_B$, \begin{equation} \label{Bogmass} m_B^2 \equiv \lambda {\tilde{C}\over 2V}~, \end{equation} appears. Diagonality of the one-particle part $:H^{(2)}:$ of the reordered Hamiltonian (\ref{hfb}) demands that we put the quasiparticle energy $\tilde{\omega}(p)$ in Eq. (\ref{phipi}) to \begin{equation} \tilde{\omega}(p)=\sqrt{p^2+m_B^2}~. \end{equation} The effective Hamiltonian depends (through $\epsilon_0$ and $m_B$) on the free parameter $\tilde{C}$. Using Eq. (\ref{Bogmass}) we can eliminate the parameter $\tilde{C}$ >from the expression for the condensate energy density $\epsilon_0$ in Eq. (\ref{Condene}) to obtain \begin{equation} \epsilon_0 =\frac{m_B^4 }{2 \lambda}~, \end{equation} which has the same nonanalytic dependence on the coupling constant as that of the Higgs mechanism of mass generation. Obviously, the Bogoliubov mechanism of spontaneous mass generation presented above differs from the Higgs mechanism by the representations of the vacuum and the interaction of the quasiparticles. The Higgs mechanism corresponds to the coherent vacuum representation, while the Bogoliubov one - to the squeezed vacuum, see Appendix \ref{app:bogmod}. The introduction of the Bogoliubov condensate is related to the Wick reordering procedure with respect to a new Fock space. \section{Bogoliubov condensate in QCD} \label{sec:qcd} After the introductory generalization of Bogoliubov condensation for superfluid $^4$He to massless $\lambda\phi^4$ theory above it is attractive to suppose that also the gluon vacuum of QCD can be considered as a homogeneous colourless condensate of gluon pairs. First steps in this direction where undertaken in Celenza and Shakin \cite{celenza}. The corresponding treatment in QCD is far more complicated than in $\lambda\phi^4$ due to the fact that QCD is a gauge theory with unphysical degrees of freedom in the Lagrangian which have to be eliminated before quantization. According to Dirac \cite{dirac}, only the spatial components of the gauge fields are dynamical and have to be quantized. The time components obey constraint equations (Gauss laws) and have to be eliminated. As in the simpler case of massless $\lambda\phi^4$ theory the squeezed condensate is described by the procedure of Wick reordering with a free parameter $\tilde{C}$ and leads both to a vacuum energy and to a mass term for the field. In QCD, however, the presence of a condensate and the corresponding generation of a mass term for the gauge field leads to spontaneous gauge symmetry breaking, since the gauge invariance of the Hamiltonian is not shared by the vacuum. The problem of the appearance of a constituent gluon mass und the corresponding resurrection of the longitudinal component of the massive quasigluons is solved by using a projection scheme for the elimination of the unphysical components of the gluon vector field \cite{dirac,perv1,perv2,fj,kp} instead of gauge-fixing. The projectors are obtained by formally solving the Gauss law constraint and appear in the kinetic energy term of a reduced gauge invariant QCD Hamiltonian. We show that these projectors are destroyed by the interaction of gluons with the squeezed vacuum so that the constituent gluon mass appears together with the necessary longitudinal components. The presence of a squeezed condensate leads to spontaneous gauge symmetry breaking: The gauge invariance of the QCD Hamiltonian is not shared by the vacuum. Finally we fix the free parameter $\tilde{C}$ of our squeezed vacuum by estimating a value for the quasigluon mass from the $\eta'-\eta$ mass difference and comparing the corresponding condensate energy density to the well known value obtained by Shifman et al. \cite{zakharov}. We shall see that they are in good agreement. \subsection{QCD Hamiltonian and Gauss law} We start from the QCD Lagrangian \begin{eqnarray} \label{1} {\cal L}(A) = -\frac{1}{4} F_{\mu\nu}^a F^{\mu\nu a}~, \end{eqnarray} where $F_{\mu\nu}^a$ is the field strength tensor \begin{eqnarray} F^a_{\mu\nu} = \partial_{\mu} A^a_{\nu} - \partial_{\nu} A^a_{\mu} + g f^{abc} A^b_{\mu} A^c_{\nu}~. \end{eqnarray} In the following, we use the notation \begin{eqnarray} {A}_{\mu} = g \frac{A^a_{\mu} \lambda^a}{2i}~, \end{eqnarray} where $ g$ is the coupling constant. Due to the gauge invariance \begin{eqnarray} \label{GI} {A}^v_{\mu} = v({A}_{\mu} + \partial_{\mu})v^{-1}~, \end{eqnarray} \begin{eqnarray} {\cal L} (A^v) = {\cal L} (A)~, \end{eqnarray} this classical Lagrangian contains only $3(N_c^2-1)$ degrees of freedom instead of the $4(N_c^2-1)$ components of the ${A}$ field ($N_c$ is the number of colours). For the construction of the Hamiltonian one usually introduces chromoelectric and chromomagnetic fields \begin{eqnarray} E^a_i&=&\dot{A}^a_i - D^{ab}_i({\bf A}) A_0^b~,\\ B^a_i({\bf A}) &=& \frac{1}{2}~\epsilon_{ijk} F_{jk}^a = \epsilon_{ijk}D^{ab}_j({\bf A}) A^b_k~, \end{eqnarray} with $\dot{A}_i = {\partial_0} A_i $ and the covariant derivative \begin{eqnarray} D^{ab}_i ({\bf A}) = \delta^{ab} \partial_i + g~{f^{acb}} A^c_i~. \end{eqnarray} The magnetic field satisfies the Bianchi identity \begin{eqnarray} \label{BI} D^{ab}_i ({\bf A}) B^b_i({\bf A}) \equiv 0 ~, \end{eqnarray} which can be interpreted as a generalized transversality of the magnetic field. In order to construct the Hamiltonian, we have to find the canonical momenta. We see that the Lagrangian (\ref{1}) does not contain time derivatives of the zero components of the gluon fields. The corresponding Euler-Lagrange equations are therefore constraints (the Gauss laws): \begin{eqnarray} \label{Gaulaw} D^2_{ab} ({\bf A}) A^b_0 = D^{ab}_i ({\bf A}) \dot{A}^b_i~, \end{eqnarray} where $D^2_{ab} = D^{ac}_i D^{cb}_i$. In terms of the electric field the Gauss laws (\ref{Gaulaw}) read \begin{equation} \label{clGau} G^a({\bf A},{\bf E})\equiv D^{ab}_i({\bf A})E^b_i=0~. \end{equation} The canonical momenta to the spatial fields $A_i^a$ are the electric fields: \begin{equation} {\delta{\cal L}\over \delta \dot{A}^a_i}=E_i^a ~,~~i=1,2,3~. \end{equation} The Hamiltonian can now be written as \begin{equation} \label{clHam} H({\bf A},{\bf E})=\int d^3x {1\over 2}\left[{E_i^a}^2+{B_i^a}^2\right]~. \end{equation} In the classical theory we thus have the Hamiltonian (\ref{clHam}) together with the Gauss constraint (\ref{clGau}). \subsection{Quantization} In order to quantize the theory one could write the Hamiltonian completely in terms of gauge invariant variables and their canonical conjugate momenta and then impose the canonical commutation relations only on these gauge invariant variables. Different to the case of QED, this leads to inconsistencies in QCD, as shown in detail in Appendix B. A more successful alternative way is Dirac quantization \cite{dirac}, where one imposes the canonical commutation relations on the original $A_i$ and $E_i$: \begin{equation} \label{QCDCCR} [E_i^a({\bf x}),A_j^b({\bf x}')]= i\delta^{ab}\delta_{ij}\delta({\bf x}-{\bf x}')~. \end{equation} Both the Hamiltonian $H$ in (\ref{clHam}) and the Gauss function $G^a$ in (\ref{clGau}) then become operators satisfying \begin{eqnarray} \label{GIC} [H,G^{a}({\bf x})] &=& 0~, \\ \quad [ G^{a}({\bf x}),G^{b}({\bf x'})] &=& if^{abc}G^c({\bf x})\delta({\bf x}-{\bf x}'). \end{eqnarray} Since $G^a$ can be interpreted as the infinitesimal generator for gauge transformations these two commutation relations express the gauge invariance of the Hamiltonian (under small gauge transformations). Note that the Hamilton operator obtained from the classical Hamiltonian (\ref{clHam}) with the cartesian fields $E_i$ and $A_i$ has the correct operator ordering \cite{ChrLee}. As pointed out by Jackiw \cite{RJ}, the Gauss law $G^a=0$ cannot be taken as an operator equation since it would lead to inconsistency with the Dirac commutation relations (\ref{QCDCCR}). Jackiw then suggested that the Gauss law should be implemented by demanding that a physical state satisfies the Schr\"odinger equation (with energy eigenvalue $\cal E$) and is annihilated by the Gauss law operator \begin{eqnarray} H({\bf A},{\bf E})|\Phi>&=&{\cal E} |\Phi>~,\\ \label{GauCon} G^a({\bf A},{\bf E})|\Phi>&=& 0~. \end{eqnarray} The second equation is the condition of gauge invariance of the physical states. The Gauss law constraint (\ref{GauCon}) can then at least in principle be implemented by use of unitary transformations \cite{GJ} and is still under lively discussion \cite{Lenz}. The resulting kinetic term in the Hamiltonian is very complicated. The requirement of gauge invariance of the physical states expressed by (\ref{GauCon}), however, is too restrictive. It does not allow for the possibility of spontaneous breaking of the gauge symmetry in analogy to spontaneous symmetry breaking in the Higgs model and spontaneous chiral symmetry breaking. If the gauge symmetry is broken spontaneously, the gauge invariance of the Hamiltonian is not shared by the vacuum. We can arrive at a gauge invariant Hamiltonian without demanding gauge invariance of the physical states, especially of the vacuum \begin{equation} G^{a}({\bf E},{\bf A})|0_B>\neq 0~, \end{equation} by starting from a gauge invariant reduced classical Hamiltonian. This is achieved by a projection method described in the following. Using the formal solution of the Gauss equations (\ref{Gaulaw}) \begin{equation} \label{10} A^a_0[{\bf A}] = \frac{1}{D^2_{ab}({\bf A})} D^{bc}_i ({\bf A})\dot{A}_i^c~, \end{equation} the electric field can be written as \begin{equation} \label{transvE} E^a_i=\Pi_{ij}^{ab}({\bf A})\dot{A}_j^b~, \end{equation} with the projection operator \begin{equation} \label{pa} \Pi_{ij}^{ab}({\bf A}) = \delta_{ij}\delta^{ab}-D^{ac}_i({\bf A}){1\over D^2_{cd}({\bf A})} D^{db}_j({\bf A})~. \end{equation} We assume that zero modes of the differential operator $D^2_{ab}(A)$ are absent. Consideration of zero modes is under current investigation \cite{kp}. In the case of $A=0$, this projection operator reduces to the transverse one $\Pi_{ij}^{ab}({\bf A}=0) =\delta^{ab} \delta^T_{ij}\equiv\delta^{ab} \left(\delta_{ij}-\partial_i\partial_j/\partial^2\right)$. The gauge invariant reduced Lagrangian can be written as \begin{equation} \label{redLag} {\cal L}^{Red}({\bf A})= {1\over 2}\left[\left(\Pi_{ij}^{ab}({\bf A})\dot{A}_j^b\right)^2- {B_i^a}^2({\bf A})\right]~. \end{equation} Note that we still have \begin{equation} {\delta{\cal L}^{Red}\over \delta \dot{A}^a_i}=E_i^a~, ~~i=1,2,3~. \end{equation} Due to the property $\Pi^2=\Pi$ of the projection operator the gauge invariant reduced Hamiltonian can be written in the form \begin{eqnarray} \label{redHam} H^{Red}({\bf A},{\bf E})= \int d^3x {1\over 2}\left[E_i^{a}\Pi^{ab}_{ij}({\bf A})E_j^b+{B_i^{a}}^2({\bf A}) \right]~. \end{eqnarray} The non-Abelian projection operator has been inserted between the cartesian electric fields $E_i$, which as variables of the Hamiltonian lost their transversality property. The above form (\ref{redHam}) is only one of many possible forms $E\Pi E$, $(\Pi E)^2$, $(\Pi E)\Pi (\Pi E)$,... Whereas in QED they are equivalent due to the property $\Pi^2=\Pi$ and the possibility to perform partial integrations, in QCD they are inequivalent due to the presence of the $A$ field in the covariant derivatives and lead to different operator orderings of $E$ and $A$ after quantization. The simplest choice $E\Pi E$ in (\ref{redHam}) will be correct at least for our investigation of a squeezed homogeneous condensate, as discussed in the next paragraph. Although the form (\ref{redHam}) is gauge invariant classically, we did not yet succeed in showing explicitly, that the corresponding Hamilton operator satisfies (\ref{GIC}). Thus in our treatment the role of gauge fixing is played by the projection operator (\ref{pa}), for deatils see \cite{perv1,perv2,kp}. The nonabelian chromomagnetic field projects onto the generalized transverse component of the $A$ field quite analogous to the form $E_i^a \Pi_{ij}^{ab}(A) E_j^b$ for the chromoelectric field. \subsection{Spectrum of quasigluon excitations} \label{ssec:qge} We shall consider the squeezed vacuum containing a colourless homogeneous condensate of gluon pairs for the Hamiltonian \begin{equation} \label{QCDHAM} :H({\bf A},{\bf E}): = \int {\rm d}^3 x ~ \frac{1}{2} \left[ :E_i^a \Pi_{ij}^{ab}({\bf A}) E_j^b: + :{B^a_i}^2({\bf A}): \right]~. \end{equation} We have introduced normal ordering with respect to creation $a_p^+$ and annihilation operators $a_p$ defined with respect to some open $\tilde{\omega}(p)$ in close analogy to our definitions (\ref{FouTra}) and (\ref{phipi}) introduced in Section II for the $\lambda\varphi^4$ model. A four-gluon interaction term occurs in the $AAAA$ term of the magnetic part ${B^{a}_i}^2({\bf A})$ and in the kinetic term $E_i^a \Pi_{ij}^{ab} E_j^b$. In analogy to the $\lambda \varphi^4$ model we perform Wick reordering to the new squeezed vacuum and consider a homogeneous and colourless condensate ($f_{p\neq0}=0, f_0\neq 0$) with the contraction \begin{eqnarray} \label{aa} <0 \,|\,A_i^a(p_1)A_j^b(p_2) \,|\,0>&=& <0_B\,|\,(A^B)_i^a(p_1) (A^B)_j^b(p_2)\,|\,0_B> = \delta_{ij} \delta^{ab} \delta_{p_1,0}\delta_{p_2,0} C~~, \end{eqnarray} where $A({\bf p}),E({\bf p})$ are the Fourier transforms of $A({\bf x}),E({\bf x})$ in analogy to (\ref{bc}), and the reordering formula for zero momentum gluon fields \begin{equation} \label{apa} :A_k^a(p=0) A_l^b(p=0): = ::{(A^B)}^a_k(p=0){(A^B)}^b_l(p=0):: {\rm e}^{2f_0} + \tilde{C} \delta^{ab} \delta_{k,l}~, \end{equation} is in analogy to Eq. (\ref{bc}), where $\tilde{C}= C \left (\mbox{\large{e}}^{2f_{0}} -1 \right)$. The corresponding reordering formula for the quartic term is calculated in Appendix C. The Wick reordering of the magnetic part of the Hamiltonian (\ref{17}), \begin{eqnarray} \frac{1}{2}\int {\rm d}^3 x ~:{B_i^a}^2({\bf A}): &=& \frac{1}{2}\int {\rm d}^3 x ~\left\{:(\partial_j A_k^a) \delta^T_{kl} (\partial_jA_l^a): +2 g f^{abc} :(\partial_j A_k^a) A_j^b A_k^c: +\frac{1}{2} g^2 f^{abc} f^{ade} : A_j^bA_k^cA_j^dA_k^e : \right\}~, \end{eqnarray} leads to the result (see Appendix \ref{app:wick}) \begin{eqnarray} \label{b2} \frac{1}{2}\int {\rm d}^3 x ~: B_i^{a2}({\bf A}): &=& g^2\frac{3}{2} \frac{N_c}{V} (N_c^2 - 1 ) {(\tilde{C})}^2 +\frac{1}{2} \sum_{p\neq 0} \left[(p^2 + 2 g^2 N_c \tilde{C}/V) \delta_{ij} - p_i p_j \right] : A_i^a(p) A_j^a(-p) : \nonumber\\&&+\frac{1}{4 V} g^2 \sum_{p_1 \dots p_4\neq 0}f^{abc} f^{ade} :A_j^b(p_1)A_k^c(p_2)A_j^d(p_3)A_k^e(p_4) : \delta_{p_1+p_2+p_3+p_4,0} ~. \end{eqnarray} The first term corresponds to the conventional definition of the gluon condensate \cite{zakharov} \begin{eqnarray} \label{26} G^2 &=& \frac{ g^2}{\pi^2} < \frac{1}{4} F^a_{\mu\nu} F^{\mu\nu a} > = \frac{g^2}{2 \pi^2} < B^2_i > \\ \nonumber &=& \frac{N^2_c - 1}{2\pi^2 3N_c} (3N_c g^2 \tilde{C}/V)^2~. \end{eqnarray} The second term in Eq.(\ref{b2}) includes the mass of the quasigluons which have both transverse and longitudinal parts, as the operation of the reordering destroys the projection properties of the non-Abelian magnetic field. This fact can be understood as spontaneous gauge symmetry breaking, and is the Bose-analogy of the spontaneous chiral symmetry breaking for the constituent quarks \cite{ad,lopr,pk2s}, which is also realized by the corresponding Bogoliubov transformation of the ''squeezed'' type. It is well known that a massive vector field requires for the description a number of degrees of freedom which exceeds that provided after fixation of a gauge in QCD. Note that the fixing of a gauge results in an elimination of the longitudinal components of the vector field and is inconsistent with the concept of a mass \cite{slavnov}. So, the method of projection onto gauge invariant variables \cite{dirac} used here is more adequate to the phenomenon of spontaneous gauge symmetry breaking than the conventional gauge fixing method. Similarly to the above magnetic energy the projector $\Pi_{ij}^{ab}$ in the electric energy $E_i^a \Pi_{ij}^{ab}({\bf A}) E_j^b$ in (\ref{QCDHAM}) is destroyed by the Wick reordering procedure leading to the kinetic energy contribution \begin{eqnarray} \label{ekin} :E_i^a E_j^b : <\Pi_{ij}^{ab}({\bf A})> ~. \end{eqnarray} The expression $<\Pi_{ij}^{ab} ({\bf A})>$ can be determined in the low energy limit $ p^2 \sim 0 $: \begin{eqnarray} \label{cpr} <\Pi_{ij}^{ab}({\bf A})> \big|_{p \sim 0}\simeq \delta_{ij}\delta^{ab} - <f^{agc} A^g_i \frac{1}{\sum_k f^{cme} A^m_k f^{end} A^n_k}f^{dlb} A^l_j > = \frac{2}{3} \delta_{ij}\delta^{ab} ~. \end{eqnarray} which can easily be checked by summation over colour and space indices. The Eqs. (\ref{b2}), (\ref{ekin}) and (\ref{cpr}) allow us to find the effective Hamiltonian for the quasigluon excitations in the low-energy limit: \begin{eqnarray} \label{lowHam} H_{\rm eff} ({\bf E},{\bf A}) = \frac{1}{2} (\frac{2}{3}) \int {\rm d}^3 x (E_i^2 + m_g^2 {A}_i^2 )~, \end{eqnarray} with the quasigluon mass \begin{equation} \label{mg2} m_g^2 = 3 N_c g^2 \frac{\tilde{C}}{V}~. \end{equation} The corresponding effective low energy Lagrangian is \begin{eqnarray} \label{lquglu} {\cal L}_{\rm eff} ({\bf A}) = \frac{1}{2} (\frac{2}{3}) \int {\rm d}^3 x (\dot{A}_i^2 - - m_g^2 {A}_i^2 )~. \end{eqnarray} The quasigluon mass $m_g$ in the low energy limit is determined by the vacuum expectation value using the relations (\ref{26}) and (\ref{mg2}): \begin{eqnarray} \label{mg} m_g = \sqrt{\frac{3\pi G}{2}}~. \end{eqnarray} We have shown that the Bogoliubov condensation of gluon pairs leads to a nonvanishing contraction $\tilde{C}$ of gluon fields which results in the spontaneous gauge symmetry breaking and the occurence of a gluon mass. We have obtained a new gluon mass formula for the low energy limit of QCD. In the following Section we examine consequences of the present approach to the low energy sector of QCD for the $\eta'$ mass formula. \section{Applications: Quasigluon mass and $\eta -\eta'$ mass difference } \label{sec:app} According to the Bogoliubov condensate approach, the contraction $\tilde{C}$ is a phenomenological parameter of the "squeezed" vacuum state and is directly related to the macroscopic occupation of the zero momentum quasigluon state and therefore to the gluon mass. We suggest to use this relationship for the fixation of the squared mass difference \begin{eqnarray} \label{eta} m_{\eta'}^2 - m_{\eta}^2 = \Delta {m^2_{\eta'}} = 0.616\; {\rm GeV}^2. \end{eqnarray} According to conventional approaches to the determination of the $\eta'$ mass \cite{rst}, we suppose that the mass difference (\ref{eta}) is determined by the $\eta' \rightarrow \eta'$ transition through the process of the anomalous decay of the $\eta'$ meson into the gluon condensate $B_i^a$ and a collective gluon excitation $E^a_i$. The effective Lagrangian of such a process can be derived according to Ref. \cite{rst}, see also \cite{volkov}, with the result \begin{eqnarray} \label{lag} {{\cal L}_{\eta'}} = \frac{1}{4} F^a_{\mu\nu}\tilde{F}^{a\mu\nu}~\eta' c_\pi~, \end{eqnarray} where $c_\pi=\sqrt3 \alpha_s /({\pi F_{\pi}})$, $\alpha_s = g^2/4\pi$ and $F_\pi = 93$ MeV. For the derivation of an effective Hamiltonian for this process, we use the sum of the Lagrangians (\ref{lag}) and (\ref{lquglu}), \begin{eqnarray} \label{ltot} {\cal L}_{\rm eff}(A)={1 \over 2}(\dot{A}_i)^2- c_\pi \eta'\dot{A}_i B_i + \dots~. \end{eqnarray} From (\ref{ltot}) follows \begin{eqnarray} E_i=\dot{A}_i - c_\pi \eta'B_i~, \end{eqnarray} such that the effective Hamiltonian reads \begin{eqnarray} {\cal H}_{\rm eff}&=& E_i\dot{A}_i - {\cal L}_{\rm eff}(A)\noindent \\ &=& \frac{1}{2} (E_i + c_\pi \eta'B_i)^2 + \dots~. \end{eqnarray} The effective Hamiltonian of the $\eta' \rightarrow \eta'$ transition has the form \begin{eqnarray} {{\cal H}}_{\eta'\rightarrow \eta'}= {1 \over 2}(\eta')^2 (B_i^{a}(b))^2 c_\pi^2~, \end{eqnarray} where $b$ is the constant part of the $A$ field. Thus, we have \begin{eqnarray} \label{eta2} \Delta {m^2_{\eta'}} = (\frac{3\alpha_s^2}{\pi^2 F_{\pi}^2})~< (B_i^a(b))^{2} >. \end{eqnarray} This equation together with Eqs.(\ref{26}) and (\ref{mg}) leads to a relation between the quasigluon and $\eta'$ masses: \begin{eqnarray} \label{eta3} \Delta {m^2_{\eta'}} =\frac{2}{3} \frac{\alpha_s}{\pi^3 F_{\pi}^2}~m_g^4~. \end{eqnarray} Choosing in the low-energy region $\alpha_s \simeq 1 $ we can estimate >from this formula the value of the quasigluon mass and by formula (\ref{mg}) then the corresponding value of the gluon condensate as \begin{eqnarray} \label{mg1} m_g = 0.71~ {\rm GeV}~~, \ \ \ \ G^2 = 0.011 ~{\rm GeV}^4~, \end{eqnarray} which is in the agreement with earlier estimates \cite{biro1,zakharov}. Note that the process of the decay of $\eta'$ into gluon fields by means of the Hamiltonian (\ref{lag}) is forbidden, as the vacuum expectation value from the magnetic field $< 0_B|B_i^a(b)|0_B >$ is equal to zero. \section{Conclusions} \label{sec:end} In the present paper we have considered the consequences of a squeezed vacuum for the single particle excitation spectrum in the gluon sector of QCD by applying the concept of the Bogoliubov theory of superfluidity to field theory. We have considered a squeezed homogeneous colourless condensate of zero momentum gluon pairs. The macroscopic occupation (squeezing) of the zero momentum mode has been achieved through Wick reordering of the QCD Hamiltonian and is characterized by a parameter $\tilde{C}$ which describes the magnitude of the condensate. The presence of the condensate leads to the occurrence of a gluon mass and thus to spontaneous gauge symmetry breaking, i.e. the gauge invariance of the Hamiltonian is not shared by the squeezed vacuum. Instead of eliminating unphysical degrees of freedom by fixing a gauge we use a projection operator method resulting from the formal solution of Gauss law. We show that the occurence of a condensate leads to a destruction of the projection property so that the generation of a mass is accompanied by the appearance of the necessary longitudinal component for the gauge field. We found that the quasigluon spectrum depends on the parameter $\tilde{C}$ of the squeezed representation, which yielded a relation between the quasigluon mass and the gluon condensate. We have fixed the quasigluon mass from the squared mass difference $m_{\eta'}^2 - m_{\eta}^2 = 0.616 ~ {\rm GeV}^2$ of $\eta$ and $\eta'$ and found that the corresponding value of the gluon condensate $G^2 = 0.012 {\rm GeV}^4$ then agrees well with the standard value $G^2 = 0.01 {\rm GeV}^4$ obtained by Shifman, Vainshtein and Zakharov. In this paper we have populated the zero momentum state directly with quasigluons whose dispersion relation was then determined selfconsistently by demanding diagonality of the one-particle sector of the Hamiltonian for nonzero momentum. This should be considered as a first attempt to explain the concept of the squeezed vacuum. In a more rigorous treatment the zero momentum state should first be occupied macroscopically with massless gluons using the freedom due to the infrared singularity of massless theories, and the resulting one-particle sector of the Hamiltonian should then be diagonalized by transformation to quasiparticles. This has been carried out for the much simpler $\lambda\phi^4$ theory in a separate work \cite{PaBla} which shows how in a more rigorous treatment the renormalization of both the mass and the bare coupling are included. Important extensions of the present approach include the study of small deviations from a homogeneous condensate, the inclusion of quark degrees of freedom and the generalization to finite temperatures. These issues are currently under investigation and will be reported in a forthcoming paper. \section*{Acknowledgement} We are grateful to D. Ebert, A.V. Efremov, E.A. Kuraev, H. Leutwyler, L.N. Lipatov, D.V. Shirkov, A.A. Slavnov, O.I. Zav'ialov for fruitful discussions. We thank S. Schmidt and D. Rischke for useful comments and critical reading of the manuscript. The work was supported in part by the RFFI, Grant No. 95-02-14411 and the Federal Minister for Research and Technology (BMFT) within Heisenberg-Landau Programme. Two of us (V.N.P. and M.K.V.) acknowledge the financial support provided by the Max-Planck-Gesellschaft and the hospitality of the MPG Arbeitsgruppe ''Theoretische Vielteilchenphysik'' at the Rostock University, where part of this work has been done. One of the authors (V.N.P.) acknowledges the hospitality of the International Centre for Theoretical Physics in Trieste. \begin{appendix} \section{The weakly nonideal Bose gas model} \label{app:bogmod} The Bogoliubov theory of the weakly interacting Bose gas \cite{nn} is described by the nonrelativistic Hamiltonian \begin{eqnarray} \label{w} H = \sum_p \frac{p^2}{2m}\,a^+_p a_p + \frac{U_0}{2V}\,\sum_{p_1 p_2 p_1^{'} p_2^{'}}\, a^+_{p_1} a^+_{p_2} a_{p^{'}_2} a_{p^{'}_1} \delta_{p_1+p_2, p_1^{'}+ p_2^{'}}~. \end{eqnarray} The operators $a^+_p, a_p$ are the creation and annihilation operators of bosons in the state $p$ satisfy the commutation relations \begin{eqnarray} \label{cr} [a_p,a^+_{p^{'}}]= \delta_{{pp}^{'}}, \end{eqnarray} where $p$ stands for momentum and internal quantum numbers. The coupling constant $U_0$ is defined by the scattering amplitude of slow particles, $V$ is the volume of the system. We figure out the original work of Bogoliubov \cite{nn}, for a more recent presentation of the theory of the weakly interacting Bose gas see \cite{FW}, \cite{R}. In the Bogoliubov derivation of the superfluid spectrum one can distinguish three points: \begin{enumerate} \item A macroscopic occupation of the zero momentum state ($p=0$) is assumed so that in the thermodynamic limit a finite density of the condensate \begin{eqnarray} \label{nb} n_B = {\mbox{\rm lim}}_{\rm th} \frac{N_0}{V} \neq 0 \end{eqnarray} occurs, where $N_0$ denotes the number of particles in the condensate. Therefore, the operators $a^+_0,a_0$ in the thermodynamic limit (\ref{nb}) are described as c-numbers \begin{eqnarray} \label{N} a_0 \simeq a^+_0 \simeq \sqrt{N_0}. \end{eqnarray} This description, strictly speaking, should be completed by defining a representation for the condensate which in the present work is given below. \item The next step is the expansion of the Hamiltonian (\ref{w}) around these c-numbers \begin{eqnarray} \label{exp} \sum a^+_{p_1} a^+_{p_2} a_{p^{'}_2} a_{p^{'}_1} & = & N^2_0 + N_0 \sum_{p \neq 0} (a^+_p a^+_{-p} + a_p a_{-p} + 4a^+_p a_p) + O[a_{p\neq 0}^3] ~. \end{eqnarray} Taking into account the conservation of the total number of particles $N$ and rewriting $$ N_0=N - \sum_{p\neq 0}a^+_p a_p ~~~; \left (\frac{N-N_0}{N} \ll 1 \right ), $$ in Eq. (\ref{exp}) and neglecting terms of higher than second order in the particle operators $a_{p\neq 0}, a^+_{p\neq 0} $, the Hamiltonian (\ref{w}) transforms into \begin{eqnarray} \label{ha2} H =\frac{N}{2} \nu + \sum_{p\neq 0} \left [a^+_p a_p \varepsilon _p + \frac{\nu}{2}\;(a^+_p a^+_{-p} + a_p a_{-p}) \right ] + O[a_{p\neq 0}^3], \end{eqnarray} where \begin{eqnarray} \varepsilon_p =\frac{p^2}{2m} + \nu, ~~~ \nu=U_0\frac{N}{V}. \end{eqnarray} \item The last step is the diagonalization of (\ref{ha2}) using the Bogoliubov transformation, i.e. the transition to the operators of quasiparticles $b^+_p $ and $b_p $ for $p \neq 0$ \begin{eqnarray} \label{b1} b_p &=& U^{-1} a_p U = {\rm cosh}(f_p) a_p + {\rm sinh}(f_p) a^+_{-p}, \nonumber\\ b^+_p&=& U^{-1} a^+_p U = {\rm cosh}(f_p) a^+_p + {\rm sinh}(f_p) a_{-p}, \\ \nonumber \end{eqnarray} where \begin{eqnarray} \label{b} U =\exp \left \{\sum_p \frac{f_p}{2} (a^+_pa^+_{-p}\, - \,a_p a_{-p}) \right \}~. \end{eqnarray} The $b_p$ satisfy the same commutation relations as the $a_p$. The function $f _p $ is found from the requirement of the disappearance of nondiagonal terms as \begin{eqnarray} \label{fp} f_p = \frac{1}{2} {\rm arth} \left[\frac{\nu}{\varepsilon_p}\right]~, \end{eqnarray} so that the Hamiltonian (\ref{ha2}) gets the form \begin{eqnarray} H =\frac{N}{2} \nu -{1\over 2}\sum_{p\neq 0} \left(\varepsilon_p-\omega_B(p)\right) + ~ \sum_{p\neq 0} b^+_p b_p~ \omega_B(p) +~ O[b_{p\neq 0}^3]~, \end{eqnarray} where $\omega_B$ is the spectrum of excitations in a superfluid liquid \begin{eqnarray} \label{bf} \omega^2_B(p) = \varepsilon^2_p - \nu^2 =\left (\frac{p^2}{2m}\right )^2 + \frac{p^2}{2m} \;\left ( 2U_0\; \frac{N}{V} \right ) \;\;, \end{eqnarray} which is determined by the condensate density $n_B = N_0/V \cong N/V$ and by the coupling constant $U_0$. \end{enumerate} In the low momentum region this expression describes the Landau sound and the particle excitations with energy $\left (p^2/2m\right)$ disappear. Note that the vacuum energy $E_0$ contains a divergent sum which can be renormalized by expressing it in terms of the physical scattering length $a$ instead of the bare coupling $U_0$ (see \cite{FW}, p. 318). In his paper \cite{nn}, Bogoliubov did not determine the representation of the condensate state for which Eq. (\ref{N}) is fulfilled. Usually one assumes that of the coherent state \begin{eqnarray} \label{c} |0_{\rm C} >~ = \exp\left\{\sum_p c_0(a^+_0 + a_0)\right\}~ |0>~,~ c_0 =\sqrt{N_0}~, \end{eqnarray} for which holds \begin{eqnarray} <0_{\rm C}|a_0|0_{\rm C}>~=~<0_{\rm C}|a^+_0|0_{\rm C}>~=\sqrt{N_0}~, \end{eqnarray} corresponding to Eq. (\ref{N}). However, to get the Bogoliubov result it is enough to assume the weaker condition \begin{eqnarray} \label{nn} (a^+_0)^2 \simeq a^2_0 \simeq a^+_0\; a_0 \sim N_0~, \end{eqnarray} rather than (\ref{N}). These relations are fulfilled for the representation of the condensate state which is given by the same Bogoliubov transformation as for $p \neq 0$ (\ref{b}): \begin{eqnarray} \label{s} |\,0_{B} > = U_{B}^{-1}~~|~0 >~, \end{eqnarray} where \begin{eqnarray} \label{b0} U_{B_0} =\exp \left \{ \frac{f_0}{2} (a^+_0a^+_{-0}\, - \,a_0 a_{-0}) \right \}~. \end{eqnarray} The inverse of the unitary operator (\ref{b}) defines also the transformation of the old into a new vacuum state for momenta $p \neq 0$. In quantum optics the vacuum $ |\,0_B >$ is called 'squeezed vacuum', see e.g. \cite{optics}. For the ''squeezed'' vacuum representation (\ref{s}) of the condensate we have the realization (\ref{nn}) \begin{eqnarray} \label{a2} <0_B\,|\,a^2_0\,|\,0_B> & = & <0_B\,|\,(a^+_0)^2\,|\,0_B > ~= - {\rm cosh}{f_0} ~{\rm sinh}{f_0} \;\;, \\ \nonumber <0_B\,|\,a^+_0 a_0\,|\,0_B> & = & ({\rm sinh}{f_0})^2 = N_0~, \label{a+a} \end{eqnarray} and at large $N_0$ (\ref{nn}) means that \begin{eqnarray} \label{n0} - -{\rm cosh}{f_0}~ {\rm sinh}{f_0} &\simeq & ({\rm sinh}{f_0})^2 \simeq N_0 \rightarrow \infty\;\;, \nonumber\\ f_0 \sim -\frac{1}{2}\ln{4N_0}~. \end{eqnarray} The choice of the squeezed vacuum is more prefable from the point view of a general consideration of all momenta, $p=0$ {\it and} $p \neq 0$. Together, the Bogoliubov transformation now is given by the product $UU_{B}$. \section{Gauge invariant variables} \label{app:gaugi} The unphysical components of the gluon fields are formally eliminated by the transformation to gauge invariant variables \cite{dirac,perv1} which are functionals constructed using the solution (\ref {10}) \begin{eqnarray} \label{12} {A}^I_i [{\bf A}] \equiv V({\bf A}) ({A}_i + \partial_i) V({\bf A})^{-1}~. \end{eqnarray} The matrix $V$ is defined from the equation \begin{eqnarray} \label{DefV} V({A}_0[{\bf A}] + \partial_0) V^{-1} = 0 \Rightarrow V({\bf A}) = T{\rm exp}(\int^t {A}_0[{\bf A}] dt^{'}) \end{eqnarray} (up to a stationary matrix as the time boundary condition). The invariance of these functionals under arbitrary time dependent gauge transformations $v({\bf x},t)$ \begin{equation} \label{GI2} {A}^I_i [{\bf A}^v] = V({\bf A}) v^{-1} v ({A_i} + \partial_i) v\ v^{-1} V({\bf A}) = {A}^I_i [{\bf A}]~, \end{equation} follows from the transformation properties of $A_i$ in (\ref{GI}) and of $V({\bf A})$: \begin{eqnarray} V({\bf A}^v) = V({\bf A}) v^{-1}~. \end{eqnarray} which follows from (\ref{DefV}) and (\ref{GI}). As consequence of this the variables (\ref{12}) represent only $2~(N_c^2-1) $ independent degrees of freedom. They contain hidden projection operators onto generalized transverse components similar to the magnetic field which satisfies the Bianchi identity (\ref{BI}). The projection operator is contained (different to the QED case) not in the $A_i^I$ themselves, but only in their time derivatives $\dot{A}_i^I$ which satisfy the ``Bianchi type'' identities \begin{equation} \label{DAI} D^{ab}_i({\bf A}^I) \dot{A}_i^{I^b} \equiv 0~. \end{equation} In the terms of the functionals (\ref{12}) the Lagrangian (\ref{1}) takes the form \begin{equation} \label{17} {\cal L}^{Red} ({\bf A}^I) = \frac{1}{2}~ \left [\dot{A}^{I^a2}_i - - B^{a2}_i({\bf A}^I) \right ]~. \end{equation} The canonical momenta to the spatial fields $A^{Ia}_i$ are \begin{equation} {E^I_i}^a \equiv\frac{\delta {\cal L}}{\delta \dot{A^I_i}^a}= \dot{A^I_i}^a~. \end{equation} The Hamiltonian becomes \begin{equation} H^{Red}({\bf A}^I,{\bf E}^I)= \int d^3x {1\over 2}\left[E_i^{Ia2}+B_i^a({\bf A}^I)^2 \right]~. \end{equation} It follows from (\ref{DAI}) that the electric fields ${E^I}_i^a $ satisfy the Gauss constraint \begin{equation} D^{ab}_i({\bf A}^I) {E^I}^b_i = 0~. \end{equation} Like the Bianci identity (\ref{BI}) for the magnetic field, this shows the generalized transversality of the invariant electric fields ${E^I}^a_i$. In order to quantize the theory one could then like in QED impose the following canonical commutation relations on the physical variables $A^I$ and $E^I$: \begin{equation} [E_i^{Ia}({\bf x}),A_j^{Ib}({\bf x}')]= i\delta^{ab}\delta_{ij}\delta({\bf x}-{\bf x}')~. \end{equation} In QCD, however, this leads to a contradiction when applying the covariant derivative on it. Instead one has to impose canonical commutation relations directly on the three cartesian fields $A_i$ and $E_i$ and write both $E^I$ and $A^I$ as functionals of $E$ and $A$. Whereas the form of $A^I[{\bf A}]$ is known by construction (\ref{12}), the functional form of $E^I[{\bf E},{\bf A}]$ in terms of $E$ and $A$ has not been found yet, but is subject of intensive research \cite{Kved}. However, even if this problem is solved there remains still the question of correct ordering of the operators $A$ and $E$. \section{Wick reordering of gluon fields} \label{app:wick} In this appendix we perform the Wick reordering of the magnetic part of the Hamiltonian (\ref{17}), \begin{eqnarray} \label{b2n} \frac{1}{2}\int {\rm d}^3 x ~:{B_i^a}^2({\bf A}): &=& \frac{1}{2}\int {\rm d}^3 x ~\left\{:(\partial_j A_k^a) \delta^T_{kl} (\partial_jA_l^a): +2 g f^{abc} :(\partial_j A_k^a) A_j^b A_k^c: +\frac{1}{2} g^2 f^{abc} f^{ade} : A_j^bA_k^cA_j^dA_k^e : \right\}~, \end{eqnarray} which reads in momentum space \begin{eqnarray} \label{b2p} \frac{1}{2}\int {\rm d}^3 x ~ : {B_i^a}^2({\bf A}) : &=& \frac{1}{2}\bigg\{\sum_{p} : A_k^a(p)\left[p_kp_l - p^2 \delta_{kl}\right]A_l^a(-p) :\nonumber\\ & &+2 g^2 \frac{1}{\sqrt{V}} f^{abc} \sum_{p_1 p_2} i(p_1+p_2)_j : A_k^a (p_1+p_2) A_j^b(-p_1) A_k^c(-p_2) :\nonumber\\ & & + \frac{1}{2} g^2 \frac{1}{V}f^{abc} f^{ade} \sum_{p_1 \dots p_4} :A_j^b(p_1)A_k^c(p_2) A_j^d(p_3) A_k^e(p_4) : \delta_{p_1+p_2+p_3+p_4,0}\bigg\}~. \end{eqnarray} The reodering of the first two terms on the r.h.s. of Eq. (\ref{b2p}) gives no extra contraction contribution. For the Wick reordering of the third term we write \begin{eqnarray} \label{a4} f^{abc} f^{ade} : A_j^b(p_1)A_k^c(p_2)A_j^d(p_3)A_k^e(p_4) : &=& f^{abc} f^{ade} A_j^b(p_1)A_k^c(p_2)A_j^d(p_3)A_k^e(p_4)\nonumber\\ &&-f^{abc} f^{ade} <A_j^b (p_1)A_j^d(p_3)> A_k^c(p_2) A_k^e(p_4)\nonumber\\ && - -f^{abc} f^{ade} A_j^b(p_1) A_j^d(p_3) <A_k^c(p_2) A_k^e(p_4)>\nonumber\\ &&-f^{abc} f^{ade} <A_j^b(p_1) A_k^c(p_2)> A_j^d(p_3) A_k^e(p_4)\nonumber\\ && - -f^{abc} f^{ade} A_j^b(p_1) A_k^e(p_4) <A_k^c(p_2) A_j^d(p_3)> \nonumber\\ &&+f^{abc} f^{ade} <A_j^b(p_1) A_j^d(p_3)> <A_k^c(p_2) A_k^e(p_4)>\nonumber\\ && +f^{abc} f^{ade} <A_j^b(p_1) A_k^c(p_2)> <A_j^d(p_3) A_k^e(p_4)>~. \end{eqnarray} Using formula (\ref{aa}) and the identities \begin{eqnarray} \sum_{abc} f^{abc} ~f^{abc} &=& N_c (N_c^2 -1)~,\nonumber\\ \sum_{ab} f^{abc} ~f^{abe} &=& N_c \delta_{ce}~,\nonumber \end{eqnarray} in Eq. (\ref{a4}) we thus have \begin{eqnarray} \label{25} \sum_{p_1 \dots p_4}f^{abc} f^{ade} &: & A_j^b(p_1)A_k^c(p_2)A_j^d(p_3)A_k^e(p_4) : \delta_{p_1+p_2+p_3+p_4,0} = \nonumber\\ && 6 N_c (N_c^2 - 1 ) C^2 - - 4 N_c C \sum_{p} A_i^a(p) A_i^a(-p) \nonumber\\ && + \sum_{p_1 \dots p_4}f^{abc} f^{ade} A_j^b(p_1)A_k^c(p_2)A_j^d(p_3)A_k^e(p_4) \delta_{p_1+p_2+p_3+p_4,0}\nonumber\\ &=& 6 N_c (N_c^2 - 1 ) {(C^B)}^2 +4 N_c C^B \sum_{p} ::A_i^a(p) A_i^a(-p):: {\rm e}^{f_p +f_{-p}} \nonumber\\&&+ \sum_{p_1 \dots p_4}f^{abc} f^{ade} ::A_j^b(p_1)A_k^c(p_2)A_j^d(p_3)A_k^e(p_4) :: {\rm e}^{f_{p_1} +f_{p_2} +f_{p_3} +f_{p_4} } \delta_{p_1+p_2+p_3+p_4,0} ~. \end{eqnarray} Putting all together, \begin{eqnarray} \label{b2r} \frac{1}{2}\int {\rm d}^3 x ~: {B_i^a}^2({\bf A}): &=& g^2\frac{3}{2} \frac{N_c}{V} (N_c^2 - 1 ) {(C^B)}^2 +\frac{1}{2} \sum_{p} \left[(p^2 + 2 g^2 N_c C^B/V) \delta_{ij} - p_i p_j \right] :: A_i^a(p) A_j^a(-p) :: {\rm e}^{f_p +f_{-p}} \nonumber\\&&+\frac{1}{4 V} g^2 \sum_{p_1 \dots p_4}f^{abc} f^{ade} ::A_j^b(p_1)A_k^c(p_2)A_j^d(p_3)A_k^e(p_4) :: {\rm e}^{f_{p_1} +f_{p_2} +f_{p_3} +f_{p_4} } \delta_{p_1+p_2+p_3+p_4,0} ~. \end{eqnarray} This proves the result of Eq. (\ref{b2}) used in the main text. \end{appendix} \newpage
\section{Introduction} It is believed that in two-dimensions (2D) even small disorder localizes all electron states.\cite{lee85} First hint for the absence of the mobility threshold in 2D came from the calculation of the weak localization correction, $\delta\sigma(\omega)$, to the conductivity at finite frequency $\omega$. It was shown\cite{gor79} that in the limit $k_Fl\gg 1$ one has $\delta\sigma(\omega)/\sigma_0={1\over k_F l}\ln|\omega\tau|$, where $k_F$ is the Fermi momentum, $l$ is the mean free path, and $\tau$ is the elastic scattering time; $\sigma_0=(e^2/h)k_Fl$ is the Drude conductivity. The logarithmic singularity in $\delta\sigma(\omega)$ indicates that the zero-temperature ($T=0$) conductivity is metallic only if $\omega\gg \omega_0$, where \begin{eqnarray}\label{1} \omega_0\sim {1\over\tau}e^{-\pi k_F l} \end{eqnarray} is the characteristic frequency which marks the crossover to the exponentially localized regime.\cite{lee85} From the finite size, $L$, correction to the zero-frequency conductivity, which is of the order of ${1\over k_F l}\ln (L/l)$, one can estimate the localization length $\xi$ as\cite{lee85} \begin{eqnarray}\label{2} \xi=l\exp\Biggl({\pi \over 2}k_F l\Biggr). \end{eqnarray} Therefore, starting from metal, one cannot describe the $T=0$ behavior of the conductivity at frequencies $\omega\sim\omega_0$. The adequate language for this region would be the language of the localized states. Within this language the conductivity originates from the transitions between the localized states induced by an external a.c. field. Let us briefly remind the corresponding derivation of $\sigma(\omega)$, carried out by Mott\cite{mott70} in the strongly localized regime. The Hamiltonian for strongly localized electrons reads \begin{eqnarray}\label{3} {\cal H}=\sum_i \epsilon_i c_i^{\dagger}c_i+ \sum_{ij} I_{ij} (c_i^{\dagger}c_j+c_j^{\dagger}c_i), \end{eqnarray} where $\epsilon_i$ is the energy of a localized state centered at ${\bf r}={\bf r}_i$ and $I_{ij}$ is the overlap integral which falls off exponentially with distance: $I_{ij}=I_0\exp(-|{\bf r}_i-{\bf r}_j|/a)$, where $a$ is the size of the wave-function of the localized electron. The general expression for the conductivity is given by the Kubo formula \begin{eqnarray}\label{4} \sigma(\omega)={ie^2\omega\over \hbar A} \sum_{\bf ij}|\langle {\bf i}|x|{\bf j} \rangle|^2 {n_{\bf i}-n_{\bf j} \over \omega + \omega_{\bf ij} + i0} \end{eqnarray} where $\langle {\bf i}|x|{\bf j} \rangle$ is the matrix element of $x$ calculated from the {\em exact} eigenstates $|{\bf i} \rangle$ and $|{\bf j} \rangle$ of the Hamiltonian (\ref{3}) with energies $E_{\bf i}$ and $E_{\bf j}$, $\hbar\omega_{\bf ij}=E_{\bf i}-E_{\bf j}$ and $n_{\bf i}$ is the occupation number of the state $|{\bf i} \rangle$. The dissipative conductivity, $\mbox{Re}\sigma(\omega)$, is determined by the pairs of states with $\omega_{\bf ij}=\omega$. At $\hbar\omega\ll I_0$ the spatial separation between bare states $|i\rangle$ and $|j\rangle$ is much larger than $a$. Then in the calculation of eigenstates of a resonant pair one should take into account the overlap $I_{ij}$ within this pair {\em only} and neglect the overlap with all the other localized states. This gives \begin{eqnarray}\label{5} |{\bf i}\rangle = {\epsilon_i -\epsilon_j\over\Gamma}|i\rangle + {2I_{ij}\over\Gamma}|j\rangle,~~~~~~ |{\bf j}\rangle = {2I_{ij}\over\Gamma}|i\rangle + {\epsilon_j-\epsilon_i \over\Gamma}|j\rangle. \end{eqnarray} The corresponging energies are \begin{eqnarray}\label{6} E_{\bf i,j}=(\epsilon_i+\epsilon_j)/2 \pm \Gamma/2,~~~~~ \Gamma=[(\epsilon_i-\epsilon_j)^2+4I_{ij}^2]^{1/2}=\hbar\omega_{\bf ij}. \end{eqnarray} Using (\ref{5}) the matrix element in (\ref{4}) takes the form \begin{eqnarray}\label{7} \langle {\bf i}|x|{\bf j}\rangle=(x_i-x_j)I_{ij}/\Gamma. \end{eqnarray} The contribution to $\mbox{Re}\sigma(\omega)$ from pairs with shoulder ${\bf r}$ can be presented as \begin{eqnarray}\label{8} \mbox{Re}\sigma(\omega,{\bf r})= {e^2\over\hbar}{\pi x^2 I^2(r)\over \hbar \omega}F(\hbar\omega,r), \end{eqnarray} where $F(\hbar\omega,r)$ is the density of pairs with shoulder ${\bf r}$ and excitation energy $\hbar\omega$ [here $I(r)=I_0e^{-r/a}$]. The density $F(\hbar\omega,r)$ is determined by the condition that the pair is singly occupied (with energies on the opposite sides from the Fermi level). Then we have \begin{eqnarray}\label{9} F(\hbar\omega,r)&=&g^2\int d\epsilon_1\int d\epsilon_2 \theta\Biggl({\hbar\omega\over 2}+{\epsilon_1+\epsilon_2\over 2}\Biggr) \theta\Biggl({\hbar\omega\over 2}-{\epsilon_1+\epsilon_2\over 2}\Biggr) \delta\Biggl[\sqrt{(\epsilon_1-\epsilon_2)^2+4I^2(r)}-\hbar\omega\Biggr] \nonumber\\ &=&{2g^2(\hbar\omega)^2\over\sqrt{(\hbar\omega)^2-4I^2(r)}}, \end{eqnarray} where $g$ is the density of states and $\theta(x)$ is the step-function. Substituting (\ref{9}) into (\ref{8}) and integrating over ${\bf r}$ we get \begin{eqnarray}\label{10} \mbox{Re}\sigma(\omega)= {e^2\over\hbar}(2\pi^2 g^2 \hbar\omega) \int_0^{\infty}{drr^3I^2(r)\over\sqrt{(\hbar\omega)^2-4I^2(r)}}. \end{eqnarray} For $\hbar\omega\ll I_0$ the main contribution to the integral comes from $r\sim r_{\omega}=a\ln(2I_0/\hbar\omega)\gg a$ and one obtains \begin{eqnarray}\label{11} \mbox{Re}\sigma(\omega)=\sqrt{2}\pi^2 {e^2\over \hbar}(g^2a\hbar^2\omega^2 r_{\omega}^3). \end{eqnarray} Although we cannot show it explicitly, we argue below that the Mott expression (\ref{11}) for $\mbox{Re}\sigma(\omega)$ is valid also for the Anderson insulator with $k_Fl\gg 1$. When applying (\ref{11}) to the Anderson insulator one should replace $a$ by $\xi$ from (\ref{2}) and substitute $g=m/2\pi\hbar^2$ ($m$ is the electron mass). However, the question remains: what is the magnitude of $I_0$ in this case? A plausible estimate can be obtained in the spirit of the Thouless picture of localization.\cite{tho77} By definition, $I_0$ represents the splitting of energy levels of two neighboring localized states (with centers at distance $\sim \xi$). The estimate for $I_0$ emerges if one equates this splitting to the mean energy spacing for localized states centered within the area $\sim \xi^2$, so that $I_0\sim 1/g\xi^2$ (see also Ref.~\onlinecite{imr93}). Note that in 1D case a similar argument leads to $I_0\sim 1/g_1\xi$, $g_1$ being the 1D density of states. Important is that in 1D the Kubo formula can be evaluated exactly\cite{ber73} resulting in the 1D version of the Mott formula, from which one can recover the above estimate for $I_0$. Such a mapping was first established by Shklovskii and Efros.\cite{shk81} Using Eq.~(\ref{10}) one can formally evaluate $\mbox{Re}\sigma(\omega)$ for $\hbar\omega > 2I_0$. In this case the main contribution to the integral comes from $r\sim a$ and we get \begin{eqnarray}\label{12} \mbox{Re}\sigma(\omega)={3\pi^2\over 4} {e^2\over \hbar}(I_0^2g^2a^4). \end{eqnarray} Certainly, the presentation of $I(r)$ in the form $I_0e^{-r/a}$ makes sense only for $r\gg a$ This means that the numerical coefficient in Eq.~(\ref{12}) is not reliable. Clearly, at large frequencies $\omega\gg\omega_0$ the conductivity of the Anderson insulator should have the Drude form. The fact that Eq.~(\ref{12}), calculated for strongly localized electrons, is also frequency independent allows us to assume that the description of a.c. transport in the Anderson insulator based on Hamiltonian (\ref{3}) is accurate within a numerical coefficient. In other words, we assume that despite the complex structure of the electron wave-functions in the Anderson insulator the energy dependence of the matrix elements calculated between these functions is still given by (\ref{7}). Eq.~(\ref{12}) provides yet another way to estimate $I_0$ in the limit $k_Fl\gg 1$. Namely, it matches $\sigma_0$ if we take $I_0\sim (k_Fl)^{1/2}/g\xi^2$. We see that the dependence of $I_0$ on $\xi$ in both estimates is the same; the extra factor $(k_Fl)^{1/2}$ presumably can be accounted for the $\ln(k_Fl)$ corrections to the exponent of $\xi$. There is also another argument in favor of the above estimate for $I_0$. The frequency dependence of $\mbox{Re}\sigma(\omega)$ in the Anderson insulator becomes strong for $\omega\ll\omega_0$, whereas in the picture of strongly localized electrons the demarcation frequency is $\omega\sim I_0/\hbar$. Equating $I_0$ to $\hbar\omega_0$ we get $I_0=k_Fl/g\xi^2$, with another extra factor $k_Fl$. The simplified description of the Anderson insulator based on the Hamiltonian (\ref{3}) allows one to include into consideration the Coulomb correlations (i.e., the correlations in the occupation numbers of the localized states caused by electron-electron interactions) using the ideas first spelled out in Refs.~\onlinecite{efr81,shk81,efr85}. This is the main goal of the present paper. We study the effect of Coulomb correlations on both $\mbox{Re}\sigma(\omega)$ and $\mbox{Im}\sigma(\omega)$. The most drastic conclusion we come to is that due to modification of $\mbox{Im}\sigma(\omega)$ by the Coulomb correlations {\em a system of localized electrons can support surface plasmons within a certain frequency range}. We also show that these plasmons cause an additional structure in the behavior of $\mbox{Re}\sigma(\omega)$ at $\omega > \omega_0 \sim I_0/\hbar$. The paper is organized as follows. In Section \ref{sec2} we analyze the polarizability of the localized system in the absence of Coulomb correlations. In Section \ref{sec3} we introduce the Coulomb correlations and find the dispersion law for the surface plasmons. In Section \ref{sec3} we study the corrections to the dispersion law due to the resonant scattering of plasmons by pairs of localized states. In Section \ref{sec4} we calculate the plasmon contribution to $\mbox{Re}\sigma(\omega)$. Section \ref{sec5} concludes the paper. \section{Polarizability in the absence of Coulomb correlations} \label{sec2} In this section we demonstrate that without Coulomb correlations the 2D Anderson insulator cannot support surface plasmon. In the framework of the linear response theory the dispersion law of a plasmon, $\omega(q)$, is determined from the condition\cite{mahan} \begin{eqnarray}\label{13} 1=v(q)\mbox{Re}{\bf P}(\omega,q), \end{eqnarray} where $v(q)$ is the Fourier component of the electron-electron interaction $v(r)$ and ${\bf P}(\omega,q)$ is the polarization operator. Within a standard approach ${\bf P}(\omega,q)$ is calculated for non-interacting electrons described by Hamiltonian (\ref{3}): \begin{eqnarray}\label{14} P(\omega,q)={1\over A} \sum_{\bf ij}|\langle {\bf i}|e^{i{\bf qr}}|{\bf j} \rangle|^2 {n_{\bf i}-n_{\bf j} \over \hbar\omega + E_{\bf i}-E_{\bf j} + i0}. \end{eqnarray} In the absence of disorder the eigenstates $|{\bf i} \rangle$ and $|{\bf j} \rangle$ are the plain waves so that the matrix element in (\ref{14}) reduces to the delta-function $\delta({\bf i-j-q})$. Then evaluating $P(\omega,q)$ and substituting it into (\ref{13}) together with 2D Coulomb interaction $v(q)=2\pi e^2/\kappa q$ ($\kappa$ is the dielectric constant) yields the surface plasmon with the well-known dispersion law \begin{eqnarray}\label{15} \omega(q)=\Biggl[{2\pi n e^2 q\over m\kappa}\Biggr]^{1/2}, \end{eqnarray} where $n$ is the 2D concentration of electrons. The plasmon mode is undamped if $q < \omega/v_F$ where $v_F=(4\pi n)^{1/2}\hbar/m$ is the Fermi velocity (for larger $q$ the Landau damping leads to a finite $\mbox{Im}P$). The latter condition can be rewritten as $q<1/2a_B$ where $a_B=\hbar^2\kappa/me^2$ is the Bohr radius. In the case of a strong disorder the eigenstates ${\bf i}$ and ${\bf j}$ in (\ref{14}) are the localized states. The polarization operator (\ref{14}) can be evaluated in a way similar to that employed in the Introduction for calculation of a.c. conductivity. For small $q$ the matrix element in (\ref{14}) can be evaluated using (\ref{7}): \begin{eqnarray}\label{16} \langle {\bf i}|e^{i{\bf qr}}|{\bf j}\rangle=i{\bf qr}I(r)/\Gamma. \end{eqnarray} Then the contribution to $P(\omega,q)$ from the pairs with the shoulder $r$ [density ${\cal P}(\omega,q,{\bf r})$ of the polarization operator] can be presented as \begin{eqnarray}\label{17} \mbox{Re}{\cal P}(\omega,q,{\bf r})=({\bf qr})^2I^2(r) \int_{2 I(r)}^{\infty}{d\Gamma\over \Gamma } {F(\Gamma,r)\over (\hbar\omega)^2-\Gamma^2}, \end{eqnarray} where $\Gamma$ and $F(\Gamma,r)$ is given by Eqs.~(\ref{6}) and (\ref{9}), respectively (the integral is understood as principal part). Substituting $F(\Gamma,r)$ into (\ref{17}) and integrating over $\Gamma$ we obtain \begin{eqnarray}\label{18} \mbox{Re}{\cal P}(\omega,q,{\bf r})&=&2g^2({\bf qr})^2I^2(r) \int_{2I(r)}^{\infty}d\Gamma {\Gamma\over \sqrt{\Gamma^2-4I^2(r)}[(\hbar\omega)^2-\Gamma^2]} \nonumber\\ &=&-{\pi g^2({\bf qr})^2I^2(r)\over \sqrt{4I^2(r)-(\hbar\omega)^2}}, {}~~~~~~~\mbox{for}~~\hbar\omega<2I(r), \nonumber\\ &=&0,~~~~~~~~~~~~~~~~~~~~~~~~~~~~\mbox{for}~~\hbar\omega>2I(r). \end{eqnarray} We see that $\mbox{Re}{\cal P}(\omega,q,{\bf r})$ is either negative or zero, so that Eq.~(\ref{13}) cannot be satisfied. In fact, the result (\ref{18}) is almost obvious. Indeed, the imaginary part of polarization operator density, $\mbox{Im}{\cal P}(\omega,q,{\bf r})$, at small $q$ differs from $\mbox{Re}\sigma(\omega,{\bf r})$ in (\ref{8}) by a factor $e^2\omega/q^2$, so it follows from (\ref{8}) and (\ref{9}) that \begin{eqnarray}\label{19} \mbox{Im}{\cal P}(\omega,q,{\bf r}) &=&-{\pi g^2({\bf qr})^2I^2(r)\over \sqrt{(\hbar\omega)^2-4I^2(r)}}, {}~~~~~~~\mbox{for}~~\hbar\omega>2I(r), \nonumber\\ &=&0,~~~~~~~~~~~~~~~~~~~~~~~~~~~~\mbox{for}~~\hbar\omega<2I(r). \end{eqnarray} Since $\mbox{Re}{\cal P}(\omega,q,{\bf r})$ and $\mbox{Im}{\cal P}(\omega,q,{\bf r})$ are connected via the Kramers-Kronig relation, the form (\ref{19}) immediately follows from (\ref{18}). \section{Coulomb correlations and surface plasmon} \label{sec3} In the previous section the polarization operator was evaluated using the pair density (\ref{9}) which was derived for non-interacting electrons. As it was first pointed out by Efros\cite{efr81}, interactions modify strongly the density of singly-occupied pairs. The underlying physics is the following. A pair can be singly-occupied even if both energy states reside below the Fermi level. The right condition for the pair to be singly-occupied is that the addition of a second electron (which interacts with the first one) is energetically unfavorable. Such a Coulomb correlations effectively enhance the density of ``soft'' pairs (i.e., the pairs with small excitation energy $\Gamma$). Our goal is to apply the latter argument, which was presented for strongly localized system, to the Anderson insulator with large $\xi$. In order to do so we will adopt two assumptions: $(i)$ The interactions do not change the localization radius $\xi$. $(ii)$ The estimate for the overlap integral, $I_0\sim 1/g\xi^2$, is unchanged in the presence of interactions. In other words, we assume that switching on the interactions leads to the Coulomb shifts of the eigenenergies but does not affect the wave functions. Note that assumptions $(i)$ and $(ii)$ contradict those made in Refs.~\onlinecite{ale94} and \onlinecite{pol93}, respectively. As we will see below, the relevant pairs would be those with shoulder $r\sim\xi$. In other words, the relevant transitions shift the position of electron by $\sim\xi$. To establish the form of the density of singly occupied pairs $F(\Gamma,r)$ with such a shoulder one can argue as follows. An isolated region of a size $\xi$ can be viewed as a small metallic granule. The transfer of an additional electron into this granule leads to the charging energy $U=e^2/2C$, where $C$ is the capacitance of a granule. In other words, the levels in the granule get shifted\cite{shk82} by an amount $\sim U$. Consider now two neighboring granules and assume that $U\gg I_0$. Due to aforementioned charging effect the highest occupied levels in the two granules typically differ by $\sim U$. Let for concreteness the highest occupied level in the first granule be higher by $U$ than in the second one. Then the sought singly occupied pair with frequency $\Gamma$ can be composed from the top occupied states in the first granule (these states should belong to the energy interval $\Gamma+U$ measured from the highest occupied level) and the empty states in the second granule. Then the density of pairs $F(\Gamma,r)$ (with $r\sim \xi$) can be estimated as $g^2(\Gamma+U)$. With the energy splitting taken into account it can be written as \begin{eqnarray}\label{20} F(\Gamma,r)= {2g^2\Gamma(\Gamma +U)\over\sqrt{\Gamma^2-4I^2(r)}}. \end{eqnarray} Then at $U=0$ we return to (\ref{9}). Certainly, our consideration, based on artificial arranging the localized states into the granules, provides only the order of magnitude estimate of $F(\Gamma,r)$. In particular, the numerical coefficient in (\ref{20}) cannot be found from such a consideration. Our choice of numerical coefficient in (\ref{20}) provides matching with a similar expression for strongly localized regime.\cite{shk81,efr85,efr85b} With the pair density (\ref{20}) we can now easily evaluate $\mbox{Re}{\cal P}(\omega,q,{\bf r})$. Calculating the integral over $\Gamma$ in (\ref{17}) yields \begin{eqnarray}\label{21} \mbox{Re}{\cal P}(\omega,q,r) &=&-{\pi g^2({\bf qr})^2I^2(r)\over \sqrt{4I^2(r)-(\hbar\omega)^2}} \Biggl\{1+{2\over\pi}{U\over\hbar\omega} \arctan\Biggl[{\hbar\omega\over\sqrt{4I^2(r)-(\hbar\omega)^2}}\Biggr]\Biggr\}, {}~~~\mbox{for}~~\hbar\omega<2I(r), \nonumber\\ &=&{U\over\hbar\omega} {2 g^2({\bf qr})^2I^2(r)\over \sqrt{(\hbar\omega)^2-4I^2(r)}} \ln\Biggl[{\sqrt{(\hbar\omega)^2-4I^2(r)}+\hbar\omega\over2I(r)}\Biggr], {}~~~~~~\mbox{for}~~\hbar\omega>2I(r). \end{eqnarray} The expression (\ref{21}) for $\mbox{Re}{\cal P}(\omega,q,r)$ can be also obtained, using the Kramers-Kronig relations, from $\mbox{Im}{\cal P}(\omega,q,r)$ which has a simple form \begin{eqnarray}\label{22} \mbox{Im}{\cal P}(\omega,q,{\bf r})= -{\pi ({\bf qr})^2I^2(r)\over (\hbar\omega)^2}F(\hbar\omega,r) &=&-{\pi g^2({\bf qr})^2I^2(r)\over \sqrt{(\hbar\omega)^2-4I^2(r)}} \Biggl(1+{U\over\hbar\omega}\Biggr),~~~\mbox{for}~~\hbar\omega>2I(r), \nonumber\\ &=& 0,~~~\mbox{for}~~\hbar\omega<2I(r). \end{eqnarray} Note that as in (\ref{19}), $\mbox{Im}{\cal P}(\omega,q,{\bf r})\neq 0$ only for $\hbar\omega>2I(r)$. We see that the enhancement in the density of pairs with small $\Gamma$ leads to a {\em positive} sign of $\mbox{Re}{\cal P}(\omega,q,r)$ for $\hbar\omega > 2I(r)$. This is our main observation. In order to obtain $\mbox{Re}P(\omega,q)$ one should integrate $\mbox{Re}{\cal P}(\omega,q,r)$ over ${\bf r}$. For $\hbar\omega > 2I_0$ the main contribution to this integral comes from $r\sim\xi$ [like in derivation of Eq.~(\ref{12})] and we get \begin{eqnarray}\label{23} \mbox{Re}P(\omega,q)={3\pi\over 4} {q^2UI_0^2g^2\xi^4\over (\hbar\omega)^2} \ln\Biggl({\hbar\omega\over I_0}\Biggr). \end{eqnarray} For generality we will assume that there is also a gate at a distance $d$ from the plane of 2D electrons. In this case the Fourier component of electron-electron interaction has the form \begin{eqnarray}\label{24} v(q)={2\pi e^2\over \kappa q}\Biggl(1-e^{-2qd}\Biggr). \end{eqnarray} If the gate is close to the electron plane, that is $qd\ll 1$, we can expand the exponent in (\ref{24}) and get $v(q)=4\pi e^2d/\kappa$. If $d\ll \xi$ then the capacitance $C$ reduces to the capacitance of two disks with area $S=\xi^2$ separated by a distance $d$, so that $C=\kappa S/4\pi d$ and consequently $U=4\pi e^2d/\kappa\xi^2$. With these $v(q)$ and $U$, after substituting (\ref{23}) into the plasmon equation (\ref{13}), we obtain \begin{eqnarray}\label{25} 1={3\pi\over 4}(q\xi)^2\Biggl({U\over\hbar\omega}\Biggr)^2 (I_0g\xi^2)^2\ln\Biggl({\hbar\omega\over I_0}\Biggr). \end{eqnarray} Certainly, the numerical coefficient in (\ref{25}) should not be taken seriously. According to the assumtion $(ii)$, $I_0\sim g\xi^2$. Then Eq.~(\ref{25}) yields the following dispersion law for the surface plasmon \begin{eqnarray}\label{26} q(\omega)={1\over \xi}\Biggl({\hbar\omega\over U}\Biggr) \ln^{-1/2}\Biggl({\hbar\omega\over I_0}\Biggr). \end{eqnarray} We see that the dispersion law is close to acoustic. Let us establish the frequency range for the surface plasmon with dispersion law (\ref{26}). The validity of expansion (\ref{16}), $q\xi\ll 1$, implies that $U\gg \hbar\omega\ln^{-1/2}(\hbar\omega/I_0)$. On the other hand, $\hbar\omega>2I_0$. Then the frequency range for plasmon is $2I_0\lesssim\hbar\omega\lesssim U$. The nesessary condition for this range to be wide is $U\gg I_0$. The ratio $U/I_0=4\pi e^2d/\xi^2 I_0$ with $I_0\sim 1/g\xi^2$ can be presented as $8\pi^2 d/a_B$, where $a_B=\hbar^2\kappa/me^2$ is the Bohr raduis. Thus, the condition $d\gg a_B$ insures that the plasmon equation (\ref{13}) has a solution within a wide frequency range. If $d<a_B$, the screening of Coulomb interaction by the gate is strong and the number of soft pairs is not sufficient to change the sign of $\mbox{Re}P(\omega,q)$. A similar condition can be obtained from the analysis of $\mbox{Im}P(\omega,q)$, which can be derived by integration of Eq.~(\ref{22}) over ${\bf r}$ \begin{eqnarray}\label{27} \mbox{Im}P(\omega,q)={3\pi^2\over 8} {q^2I_0^2g^2\xi^4\over (\hbar\omega)^2}(\hbar\omega+U) \sim {q^2\over (\hbar\omega)^2}(\hbar\omega+U). \end{eqnarray} The origin of $\mbox{Im}P(\omega,q)$ is the interaction of a plasmon with ``resonant'' pairs having excitation energy $\omega$. In fact, $\mbox{Im}P(\omega,q)$ describes the resonant scattering of a plasmon by a pair of localized states. One can introduce a mean free path, $l$, associated with such a scattering and obtain $ql\sim\ln^{1/2}(\hbar\omega/I_0)/(1+\hbar\omega/U)$. Thus, the condition $2I_0\lesssim\hbar\omega\lesssim U$ reduces to the condition $ql\gtrsim 1$. In the absense of gate [$qd\gg 1$ in Eq.~(\ref{24})] one should take $U=e^2/\kappa\xi$, so that $U/I_0\sim \xi/a_B\gg 1$. Then after a simple algebra we obtain \begin{eqnarray}\label{28} q(\omega)={1\over \xi}\Biggl({\hbar\omega\over U}\Biggr)^2 \ln^{-1}\Biggl({\hbar\omega\over I_0}\Biggr). \end{eqnarray} It can be shown that the above analysis of the validity applies in this case as well and leads to the same frequency range $2I_0\lesssim\hbar\omega\lesssim U$. Within this range we again have $q\xi\ll 1$. Thus, one should use Eq.~(\ref{26}) for $qd<1$ and Eq.~(\ref{28}) for $qd>1$. On the other hand, the magnitude of $U$ depends on the ratio $d/\xi$. Note that for $d>\xi$ one still can have $qd<1$. In this case the dispersion law is given by Eq.~(\ref{26}) with $U=e^2/\kappa\xi$. \section{Renormalization of the plasmon dispersion law} \label{sec4} In the previous section, when calculating the polarization operator, we took into account the Coulomb correlations within a pair, but neglected the effect of polarization of surrounding pairs on a given pair. On the other hand, by averaging of the polarization operator (\ref{14}) over frequencies of pairs $\Gamma$ and their shoulders ${\bf r}$ we have effectively replaced the localized system by a medium. The average polarization of this medium gave rise to a plasmon mode. Within this procedure the ``feedback'' from surrounding pairs reduces to the interaction of a given pair with plasmons. In the present section we study the renormalization of the plasmon spectrum due to this effect. Generally, the plasmon excitation is defined as a pole in the density-density correlation function, $\Pi(\omega, {\bf q},{\bf q}')$, which is related to the polarization operator ${\bf P}(\omega, {\bf q},{\bf q}')$ by the Dyson equation \begin{eqnarray}\label{29} \Pi(\omega, {\bf q},{\bf q}')= {\bf P}(\omega, {\bf q},{\bf q}')+\int {d{\bf q}_1\over (2\pi)^2} {\bf P}(\omega, {\bf q},{\bf q}_1)v(q_1)\Pi(\omega, {\bf q}_1,{\bf q}'). \end{eqnarray} Before averaging, both ${\bf P}(\omega, {\bf q},{\bf q}')$ and $\Pi(\omega, {\bf q},{\bf q}')$ depend on two momentum variables ${\bf q}$ and ${\bf q}'$. The approximation we made above reduces to replacing of ${\bf P}(\omega, {\bf q},{\bf q}')$ by its average ${\bf P}(\omega, q)$, so that the solution of (\ref{29}) takes the form \begin{eqnarray}\label{30} \Pi(\omega, q)= {{\bf P}(\omega, q) \over 1-v(q){\bf P}(\omega, q)}. \end{eqnarray} Then the pole of $\Pi(\omega,q)$ is determined by the plasmon equation (\ref{13}). As a next step, we took for ${\bf P}(\omega, q)$ its expression (\ref{14}) for non-interacting electrons, which represents a sum of polarizations of pairs, \begin{eqnarray}\label{31} P(\omega,q)={1\over A} \sum_{\bf ij}|\langle {\bf i}|e^{i{\bf qr}}|{\bf j} \rangle|^2 P_{\bf ij}(\omega),~~~~~ P_{\bf ij}(\omega)= {n_{\bf i}-n_{\bf j} \over \hbar\omega + E_{\bf i}-E_{\bf j} + i0}, \end{eqnarray} and performed the summation neglecting correlations between the pairs but with the Coulomb correlations within a pair included. The renormalized ${\bf P}_{\bf ij}(\omega)$ for a given pair can be obtained from the following procedure. The function $\Pi(\omega, q)$ has a diagrammatic presentation in a form of a series of bubbles $({\bf ij})$, corresponding to $P_{\bf ij}$, connected by the Coulomb interaction lines (see Fig.~1a). First, we arrange into a single block the sum over all combinations of bubbles which appear between two bubbles $({\bf ij})$ (see Fig.~1b). Then we replace this block by its average, so that the result can be presented as two bubbles $({\bf ij})$ connected by a plasmon propagator $\Pi(\omega, q)$ (see Fig.~1b) (there is also an extra factor $v(q)$ in each vertex). The renormalized bubble $({\bf ij})$ can be then obtained by summing up the series, consisting from the bubbles $({\bf ij})$, connected by plasmon lines (see Fig.~1c). The resulting expression for ${\bf P}_{\bf ij}(\omega)$ reads \begin{eqnarray}\label{32} {\bf P}_{\bf ij}(\omega)= {P_{\bf ij}(\omega,) \over 1-P_{\bf ij}(\omega)R_{\bf ij}(\omega)}, \end{eqnarray} with \begin{eqnarray}\label{33} R_{\bf ij}(\omega)=\int {d{\bf q}\over (2\pi)^2} |\langle {\bf i}|e^{i{\bf qr}}|{\bf j} \rangle|^2 v^2(q)\Pi(\omega, q). \end{eqnarray} Finally, replacing $P_{\bf ij}(\omega)$ in (\ref{31}) by ${\bf P}_{\bf ij}(\omega)$ we obtain \begin{eqnarray}\label{34} {\bf P}(\omega,q)={1\over A} \sum_{\bf ij}|\langle {\bf i}|e^{i{\bf qr}}|{\bf j} \rangle|^2 {n_{\bf i}-n_{\bf j} \over \hbar\omega + E_{\bf i}-E_{\bf j} - (n_{\bf i}-n_{\bf j})R_{\bf ij}(\omega)}. \end{eqnarray} The equations (\ref{30}), (\ref{33}), and (\ref{34}) form a closed system which determines $\Pi (\omega, {\bf q})$ and, correspondingly, the renormalized dispersion law of a plasmon in a self-consistent way. The approximation made in order to get the closed system [replacement of a block by a sought function $\Pi (\omega, q)$] is known as the effective-medium approximation. The analysis of the system (\ref{30},\ref{33},\ref{34}) reveals that the renormalization of the plasmon dispersion law is weak. Namely, the appearence of the term $R_{\bf ij}(\omega)$ in the denominator of (\ref{34}) has the physical meaning that a pair $({\bf ij})$ acquires a finite life-time $\tau_{\bf ij}$ due to the interaction with plasmons. We will show that this life-time is long, i.e., $1/\tau_{\bf ij}\ll \omega_{\bf ij}$. It can be readily seen that the difference between the renormalized polarization $\mbox{Re}{\bf P}(\omega,q)$ and $\mbox{Re}P(\omega,q)$ originates from resonant pairs with $(\omega -\omega_{\bf ij})\sim 1/\tau_{\bf ij}$ and $r\sim \xi$ (note that the main term, $\mbox{Re}P(\omega,q)$, is determined by the entire interval $\omega_{\bf ij}\sim \omega$). If we neglect the dependence of the matrix element in (\ref{34}) on $\omega_{\bf ij}$ then the renormalization correction to $\mbox{Re}P(\omega,q)$ would be identically zero. A finite correction results from a slight asymmetry of the matrix element within the narrow interval $(\omega -\omega_{\bf ij})\sim 1/\tau_{\bf ij}$. Then the relative magnitude of the correction is of the order of $1/\omega\tau_{\bf ij}$ with $\tau_{\bf ij}$ calculated for a pair with $\omega_{\bf ij}=\omega$. Expecting $\tau_{\bf ij}$ to be long, we can calculate it by substituting the nonrenormalized dispersion law of a plasmon into Eq.~(\ref{33}). Performing the integration we obtain \begin{eqnarray}\label{35} R_{\bf ij}(\omega)={\hbar\over\tau_{\bf ij}}&\sim&\hbar\omega \Biggl({\hbar\omega\over U}\Biggr)\Biggl({I_0\over U}\Biggr)^2 \ln^{-2}\Biggl({\hbar\omega\over I_0}\Biggr),~~~~~~~~~~\mbox{with gate,} \nonumber\\ &\sim&\hbar\omega \Biggl({\hbar\omega\over U}\Biggr)^3\Biggl({I_0\over U}\Biggr)^2 \ln^{-3}\Biggl({\hbar\omega\over I_0}\Biggr),~~~~~~~~\mbox{without gate.} \end{eqnarray} Since both ratios, $\hbar\omega/U$ and $I_0/U$ are small, the correction to the dispersion law, $\delta q(\omega)/q(\omega)\sim 1/\omega\tau_{\bf ij}$, is negligible. In the next section we will see that the corresponding renormalization of $\mbox{Re}\sigma (\omega)$ is much larger than the renormalization of the dispersion law. \section{Renormalization of the real part of the conductivity} \label{sec5} As we have seen in the previous section, the renormalization of the polarization operator results in an appearance of $i\hbar/\tau_{\bf ij}$ in the denominator of Eq.~(\ref{34}). Correspondingly, the renormalized expression (\ref{4}) for $\mbox{Re}\sigma(\omega)$ can now be rewritten as \begin{eqnarray}\label{36} \mbox{Re}\sigma(\omega)={e^2\omega\over \hbar A}\mbox{Im} \sum_{\bf ij}\Biggl[{I_{ij}(x_i-x_j)\over \hbar\omega_{\bf ij}}\Biggr]^2 {1 \over \omega + \omega_{\bf ij} + i/\tau_{\bf ij}}, \end{eqnarray} where the summation is performed over the singly occupied pairs ${\bf ij}$ and in this way the Coulomb correlations are taken into account. In the limit $\tau_{\bf ij}\rightarrow\infty$ and for $\omega >2I_0$ one should use the pair density $F(\Gamma,r)$, given by Eq.~(\ref{20}), in order to perform the summation. The result is determined by resonant pairs with $\omega=\omega_{\bf ij}=\Gamma/\hbar$: \begin{eqnarray}\label{37} \mbox{Re}\sigma(\omega)={3\pi^2\over 4} {e^2\over \hbar}(I_0g\xi^2)^2\Biggl(1+{U\over\hbar\omega}\Biggr)= \sigma_0\Biggl(1+{U\over\hbar\omega}\Biggr). \end{eqnarray} We see that within the frequency interval $2I_0<\hbar\omega<U$ the real part of the conductivity exceeds the Drude value due to Coulomb correlations. When calculating the correction to $\mbox{Re}\sigma(\omega)$ caused by the finite value of $\tau_{\bf ij}$ it is important to realize that $\hbar/\tau_{\bf ij}$ is maximal for soft pairs with small $\omega_{\bf ij}$. This is because the matrix element $\langle {\bf i}|e^{i{\bf qr}}|{\bf j} \rangle$ is proportional to $1/\omega_{\bf ij}$. In the previous Section this was not important since the correction to $\mbox{Re}P(\omega,q)$ came from the resonant pairs only. Here, however, we have $\mbox{Im}(\omega + \omega_{\bf ij} + i/\tau_{\bf ij})^{-1} \propto 1/\tau_{\bf ij}\propto 1/\omega_{\bf ij}^2$, so that the soft pairs give the main contribution to the correction $\delta\mbox{Re}\sigma(\omega)$. Assuming $\omega_{\bf ij}\ll\omega$ we can present this correction in the form \begin{eqnarray}\label{38} \mbox{Re}\sigma(\omega)&=&{e^2\over \hbar\omega A} \sum_{\bf ij}{[I_{ij}(x_i-x_j)]^2 \over (\hbar\omega_{\bf ij})^2\tau_{\bf ij}} \nonumber\\ &=&{e^2\over \hbar\omega A} \sum_{\bf ij}\Biggl[{I_{ij}(x_i-x_j) \over \hbar\omega_{\bf ij}}\Biggr]^4 \int {d{\bf q}\over (2\pi)^2} q^2v^2(q)\mbox{Im}\Pi(\omega, q), \end{eqnarray} where we have substituted $\hbar/\tau_{\bf ij}=R_{\bf ij}$ from Eq.~(\ref{33}). The sum over pairs is again evaluated with the pair density (\ref{20}) \begin{eqnarray}\label{39} {1\over A}\sum_{\bf ij}\Biggl[{I_{ij}(x_i-x_j) \over \hbar\omega_{\bf ij}}\Biggr]^4= {1\over 4}\int d{\bf r}r^4I^4(r) \int_{2I(r)}^{\infty}d\Gamma{F(\Gamma,r)\over\Gamma^4}. \end{eqnarray} The main contribution to the integral over $\Gamma$ comes from the lower limit $\Gamma\sim I(r)$. Then the integral over $r$ is again determined by $r\sim\xi$, so that the sum (\ref{39}) appears to be $\sim I_0\xi^6g^2U$. As in the previous section, the value of the integral in (\ref{39}) takes different values in the presence and in the absence of a gate. Finally we obtain \begin{eqnarray}\label{40} {\delta\mbox{Re}\sigma(\omega)\over\sigma_0} &\sim& \Biggl({\hbar\omega\over U}\Biggr)^3\Biggl({\hbar\omega\over I_0}\Biggr) \ln^{-2}\Biggl({\hbar\omega\over I_0}\Biggr),~~~~~~~~~~\mbox{with gate,} \nonumber\\ &\sim& \Biggl({\hbar\omega\over U}\Biggr)^5\Biggl({\hbar\omega\over I_0}\Biggr) \ln^{-3}\Biggl({\hbar\omega\over I_0}\Biggr),~~~~~~~~\mbox{without gate.} \end{eqnarray} Comparing (\ref{35}) to (\ref{40}) we see that both corrections are of the same order $\left(I_0/U\right)^2$ at $\hbar\omega\sim I_0$. However the correction to ${\mbox Re}\sigma$ is much bigger at $\hbar\omega\gg I_0$. This reveals a new mechanism of absorption of a.c. field: by resonant excitation of plasmons. More precisely, the field polarizes the soft pairs (with $\omega_{\bf ij}\sim I_0/\hbar$) and the induced polarization excites the plasmon waves. Therefore, the energy of a.c. field is effectively absorbed by plasmons. The rapid increase of $\delta\sigma$ with $\omega$ is caused by the number (phase volume) of plasmons which absorb the field. Note that with correction $\delta\sigma(\omega)$ the total conductivity ${\mbox Re}\sigma(\omega)$ exhibits a rather complicated behavior. For $\hbar\omega<2I_0$ the conductivity increases with $\omega$. Then it passes through a maximum at $\omega\sim I_0/\hbar$ and falls off with $\omega$ according to Eq.~(\ref{37}). However at $\hbar \omega\sim U(I_0/U)^{1/4}$ (with gate) and $\hbar \omega\sim U(I_0/U)^{1/6}$ (without gate) we have $\delta{\mbox Re}\sigma(\omega)\sim {\mbox Re}\sigma(\omega)$ and the conductivity starts rising again. On the other hand, the expression for $\delta{\mbox Re}\sigma$ was derived assuming that it is small. Therefore in the region $\delta{\mbox Re}\sigma >{\mbox Re}\sigma(\omega)$ the renormalization of $\sigma(\omega)$ by plasmons is strong. In this case one cannot calculate $\tau_{\bf ij}$ using the bare polarization operator. The full analysis of the system (\ref{30},\ref{33},\ref{34}) in this frequency range is out of the scope of the present paper. \section{Conclusion} \label{sec6} In the present paper we argue that the wave of electric field can propagate along the surface of the 2D Anderson insulator. The field originates from the density fluctuations of localized electrons. One should distinguish this wave from the usual plasmon in an ideal 2D gas with the dispersion law given by Eq. (\ref{15}): ({\em i}) the derivation of (\ref{15}) implies that $\omega\gg 1/\tau$ while we predict the existence of a plasmon at much lower frequencies $\omega\gtrsim I_0/\hbar$. The minimal frequencies $1/\tau$ and $I_0/\hbar$ differ by a factor $\exp(\pi k_Fl)$; ({\em ii}) the plasmon (\ref{15}) results from the solution of Eq. (\ref{13}) with polarization operator calculated for free electrons. Within this approximation Eq. (\ref{13}) has no solutions for {\em localized} electrons. The solution appears only if one takes into account the Coulomb correlations in the occupation numbers of the localized states. The obvious consequence of the existence of a plasmon excitation is that localized electrons ``feel'' a fluctuating electric field ${\cal E}(\omega,{\bf q})$ with the spectral density \begin{eqnarray}\label{41} \langle |{\cal E}(\omega,{\bf q})|^2\rangle= {\kappa^2\over 2\pi e^2}{q^2v^2(q)\mbox{Im}{\bf P}(\omega,q)\over [1-v(q)\mbox{Re}{\bf P}(\omega,q)]^2+[v(q)\mbox{Im}{\bf P}(\omega,q)]^2}, \end{eqnarray} where ${\bf P}(\omega,q)$ is the polarization operator (\ref{34}). The spectral density (\ref{41}) has a peak at $\omega=\omega(q)$ corresponding to the dispersion law of a plasmon, which is different with and without gate. The basic assumption of our theory is that in the presence of interactions the Anderson insulator is characterized by two energy scales: $I_0$ - the spacing between energy levels in the area $\xi^2$ and the charging energy $U$ which is the Coulomb interaction of two localized electrons separated by a distance $\sim\xi$. This energy modifies the pair density $F(\omega,r)$. Such a modification results in an enhancement\cite{shk81,efr81,efr85,efr85b} of the dissipative conductivity $\mbox{Re}\sigma(\omega)$. If we denote as $\rho(\omega)$ the matrix element of $r$ calculated for a pair with frequency $\omega$ then \begin{eqnarray}\label{42} \mbox{Re}\sigma(\omega)\propto \omega\rho^2(\omega)F(\omega, r_{\omega}). \end{eqnarray} As it was discussed in the Introduction, it is plausible to assume that for the Anderson insulator in the absence of interactions $\rho(\omega)$ behaves as $\rho(\omega)\sim r_{\omega}$ for $\omega\lesssim I_0/\hbar$ and $\rho(\omega)\sim\xi I_0/\omega$ for $\omega\gtrsim I_0/\hbar$. Since the pair density in the absence of interactions is proportional to $\omega$ we reproduce Eqs.~(\ref{11}) and (\ref{12}): \begin{mathletters}\label{43} \begin{eqnarray} \mbox{Re}\sigma(\omega)&\propto&\omega^2, {}~~~~~~~~~~~~\mbox{for}~~~\omega\lesssim I_0/\hbar,\label{43a}\\ \mbox{Re}\sigma(\omega)&=& const, {}~~~~~~~~~\mbox{for}~~~\omega\gtrsim I_0/\hbar,\label{43b} \end{eqnarray} \end{mathletters} Where we neglected a logarithmic factor in (\ref{43a}). We have assumed that the frequency dependence of $\rho(\omega)$ remains the same in the presence of interactions.\cite{note} At the same time, $F(\omega,\xi)$ is changed drastically by interactions: $F(\omega,\xi)=const$ for $\omega\lesssim U/\hbar$ and $F(\omega,\xi)\propto\omega$ for $\omega\gtrsim U/\hbar$. As a result, we get the following frequency dependence of $\mbox{Re}\sigma(\omega)$ in the presence of interactions [see also (\ref{37})]: \begin{mathletters}\label{44} \begin{eqnarray} \mbox{Re}\sigma(\omega)&\propto& \omega, {}~~~~~~~~~~~~~~\mbox{for}~~~\omega\lesssim I_0/\hbar,\label{44a}\\ \mbox{Re}\sigma(\omega)&\propto& 1/\omega, {}~~~~~~~~~~~\mbox{for}~~~I_0/\hbar\lesssim\omega\lesssim U/\hbar,\label{44b}\\ \mbox{Re}\sigma(\omega)&=& \mbox{const}, {}~~~~~~~~~\mbox{for}~~~\omega\gtrsim U/\hbar,\label{44c} \end{eqnarray} \end{mathletters} This simple analysis shows that $\mbox{Re}\sigma(\omega)$ exhibits a maximum at $\omega\sim I_0/\hbar$ (see also the end of the previous section). In fact, this maximum is intimately related to the existence of a plasmon. Indeed, for Eq. (\ref{13}) to have solutions we need a positive sign of $\mbox{Re}{\bf P}(\omega,q)$, which is equivalent to a positive $\mbox{Im}\sigma(\omega)$. At the same time, $\mbox{Im}\sigma(\omega)$ can be obtained from $\mbox{Re}\sigma(\omega)$ using the Kramers-Kronig relation. In order to trace how the modification of $\mbox{Re}\sigma(\omega)$ by Coulomb correlations leads to the change of sign of $\mbox{Im}\sigma(\omega)$, we can interpolate the frequency dependence of $\rho(\omega)$ as $\rho(\omega)\propto [(\hbar\omega)^2+4I_0^2]^{-1/2}$ and the frequency dependence of $F(\omega,\xi)$ as $F(\omega,\xi)\propto (\hbar\omega + U)$. Then we have \begin{eqnarray}\label{45} \mbox{Re}G(\omega)={\hbar\omega(\hbar\omega+U) \over (\hbar\omega)^2+4I_0^2},~~~~~~~~~~\omega>0, \end{eqnarray} where $G={\sigma\over e^2/\hbar}$ is the conductance. It can be easily seen that Eq. (\ref{45}) reproduces correctly all the limiting cases (\ref{44}) and (\ref{44}) both with and without charging effect. The imaginary part of the conductance calculated from (\ref{45}) has a simple form: \begin{eqnarray}\label{46} \mbox{Im}G(\omega)= -{\hbar\omega[2I_0-(2U/\pi)\ln(\hbar\omega/ 2I_0)] \over (\hbar\omega)^2+4I_0^2},~~~~~~~~~~\omega>0, \end{eqnarray} In Fig. 2 we have plotted $\mbox{Re}G(\omega)$ and $\mbox{Im}G(\omega)$ for different ratios $U/2I_0$. We see that in the absence of Coulomb correlations ($U=0$), $\mbox{Im}G(\omega)$ is strictly negative (in our calculation in Section \ref{sec2} it turns to zero for $\omega>I_0/\hbar$). However, at finite $U$ the change of sign occurs at $\omega=(2I_0/\hbar)\exp(\pi I_0/U)$. For $U\gg I_0$ this frequency is just $2I_0/\hbar$. As a final remark, let us outline the difference between our approach and that of Ref.~\onlinecite{fle78}. In Ref.~\onlinecite{fle78} the authors addressed electron-electron interactions in the strongly localized regime when the localization radius is much smaller than the interpair separation. In this case singly-occupied pairs can be considered as point-like dipoles. The dipole moment ${\bf p}_k$ induced by an external field ${\cal E}_0e^{-i\omega t}$ can be written as \begin{eqnarray}\label{47} p_k^{\mu}=\alpha_k^{\mu\nu}[{\cal E}_0^{\nu}+ \sum_{l\neq k}{\cal E}_k^{(l)\nu}], \end{eqnarray} where ${\cal E}_k^{(l)\nu}$ is a component $\nu$ of the electric field caused by polarization of a dipole $l$ which acts on the dipole $k$ (summation over repeating indices $\nu$ is implied). The polarizability $\alpha_k^{\mu\nu}$ of a dipole $k$ has the form \begin{eqnarray}\label{48} \alpha_k^{\mu\nu}={2e^2\over\hbar}{\rho_k^{\mu}\rho_k^{\nu}\omega_k \over \omega^2-\omega_k^2} ={e^2\over\hbar}\rho_k^{\mu}\rho_k^{\nu}\sum_{\bf ij}P_{\bf ij}, \end{eqnarray} where $\rho_k^{\mu}$ and $\omega_k$ are correspondingly the matrix element and the frequency of the dipole $k$. Here $P_{\bf ij}$ is given by Eq.~(\ref{31}) with the states ${\bf i}$ and ${\bf j}$ making up the dipole $k$. To keep the discussion simple we will assume that the polarizability is isotropic, i.e., $\alpha_k^{\mu\nu}=\alpha_k\delta_{\mu\nu}$. The field ${\cal E}_k^{(l)\nu}$ can, in turn, be expressed through ${\bf p}_k$ as \begin{eqnarray}\label{49} {\cal E}_k^{(l)\mu}=-{\partial\over\partial R^{\mu}} {{\bf p}_l{\bf R}\over R^3}\Biggl|_{{\bf R}={\bf R}_k-{\bf R}_l}, \end{eqnarray} where ${\bf R}_k$ is the position of the dipole $k$. Upon substituting (\ref{49}) into (\ref{47}) we obtain an infinite system of linear equations. It can be easily shown that iterating this system leads to the renormalization of polarizabilities of dipoles $\alpha_k$ \begin{mathletters}\label{50} \begin{eqnarray} \alpha_k={e^2\over\hbar}\rho_k^2 \sum_{\bf ij}{P_{\bf ij}\over 1-P_{\bf ij}\Sigma_k} ={2e^2\over\hbar}{\rho_k^2(\omega_k+\Sigma_k) \over \omega^2-(\omega_k+\Sigma_k)^2},\\ \Sigma_k=-\sum_l \alpha_l \Biggl({\partial\over\partial R^{\mu}}{R^{\nu}\over R^3}\Biggr) \Biggl({\partial\over\partial R^{\nu}}{R^{\mu}\over R^3}\Biggr) \Biggl|_{{\bf R}={\bf R}_k-{\bf R}_l}+\cdots. \end{eqnarray} \end{mathletters} Here $\Sigma_k$ is the self-energy. Fleishman and Anderson\cite{fle78} anylized this self-energy using the arguments similar to those put forward by Anderson\cite{and58} when he demonstrated the existence of localization transition for eigenfunctions of the Schroedinger equation with disorder. They argued that $\mbox{Im}\Sigma$ takes a finite value with non-zero probability. This means that in the absence of an external field the system \begin{eqnarray}\label{51} {\bf p}_k+\alpha_k\nabla\sum_{l\neq k} {{\bf p}_l{\bf R}\over R^3}\Biggl|_{{\bf R}={\bf R}_k-{\bf R}_l}=0, \end{eqnarray} has delocalized solutions.\cite{lev90} In other words, by analogy to the Schoedinger equation in the tight-binding approximation, the eigenstates $\{{\bf p}_k\}$ of the system ({\ref{51}) extend throughout the entire volume. Fleishman and Anderson\cite{fle78} considered a three-dimensional system and neglected the Coulomb correlations. Note that if one rewrites the system (\ref{51}) in the momentum representation and averages it by factorizing the average of the product $\alpha{\bf p}$ (mean-field) then one arrives at the plasmon equation (\ref{13}). As it was demonstrated in Section \ref{sec2}, this equation has no propagating solutions in the absence of Coulomb correlations. Our central point is that {\em with} the Coulomb correlations the delocalized solution exists even at the mean-field level. This solution, that is surface plasmon, is specific for the two-dimensional system and is characterized by the dispersion law $\omega(q)$. An interesting question that could be addressed within the same approach is how the Coulomb correlations modify the a.c. Hall conductivity $\sigma_{xy}(\omega)$ of the Anderson insulator.\cite{imr93,zha92} \acknowledgments The authors are indebted to B.~I. Shklovskii for numerous discussions. In fact, Sections \ref{sec4}-\ref{sec6} have emerged from his hot contesting of our basic assumptions.
\section*{Appendix} \setcounter{equation}{0} \defA\arabic{equation}{A\arabic{equation}} It is convenient to define the mass dependent variables $a:=2+\sqrt\xi$, $b:=2-\sqrt\xi$ and $w:=\sqrt{(1-\sqrt\xi)/(1+\sqrt\xi)}$. The integrals $t_1,\ldots,t_{12}$ appearing in Eqs.~(\ref{eqn13})--(\ref{eqn14}) are then given by \begin{eqnarray} t_1&:=&\ln\left(\frac{2\xi\sqrt\xi}{b^2(1+\sqrt\xi)}\right),\quad t_2\ :=\ \ln\left(\frac{2\sqrt\xi}{1+\sqrt\xi}\right)\quad \Rightarrow\quad t_1-t_2\ =\ \ln\left(\frac\xi{b^2}\right)\\ t_3&:=&\ln\left(\frac{1+v}{1-v}\right)\\ t_4&:=&\mbox{Li}_2(w)-\mbox{Li}_2(-w)+\mbox{Li}_2(\frac abw)-\mbox{Li}_2(-\frac abw)\\ t_5&:=&\frac12\ln\left(\frac{\sqrt\xi(2+\sqrt\xi)}{4(1+\sqrt\xi)}\right)\ln\left(\frac{1+v}{1-v}\right) +\mbox{Li}_2\left(\frac{2\sqrt\xi}{a(1+w)}\right) -\mbox{Li}_2\left(\frac{2\sqrt\xi}{a(1-w)}\right)\,+\nonumber\\&& +\mbox{Li}_2\left(\frac{1+w}2\right)-\mbox{Li}_2\left(\frac{1-w}2\right) +\mbox{Li}_2\left(\frac{a(1+w)}4\right)-\mbox{Li}_2\left(\frac{a(1-w)}4\right)\\ t_6&:=&\ln^2(1+w)+\ln^2(1-w)+\ln\left(\frac{2+\sqrt\xi}8\right)\ln(1-w^2) \,+\nonumber\\&& +\mbox{Li}_2\left(\frac{2\sqrt\xi}{a(1+w)}\right) +\mbox{Li}_2\left(\frac{2\sqrt\xi}{a(1-w)}\right) -2\mbox{Li}_2\left(\frac{2\sqrt\xi}a\right)\,+\nonumber\\&& +\mbox{Li}_2\left(\frac{1+w}2\right)+\mbox{Li}_2\left(\frac{1-w}2\right) -2\mbox{Li}_2\left(\frac12\right)\,+\nonumber\\&& +\mbox{Li}_2\left(\frac{a(1+w)}4\right)+\mbox{Li}_2\left(\frac{a(1-w)}4\right) -2\mbox{Li}_2\left(\frac a4\right)\\ t_7&:=&2\ln\left(\frac{1-\xi}{2\xi}\right)\ln\left(\frac{1+v}{1-v}\right) -\mbox{Li}_2\left(\frac{2v}{(1+v)^2}\right) +\mbox{Li}_2\left(-\frac{2v}{(1-v)^2}\right)\,+\nonumber\\&& -\frac12\mbox{Li}_2\left(-\left(\frac{1+v}{1-v}\right)^2\right) +\frac12\mbox{Li}_2\left(-\left(\frac{1-v}{1+v}\right)^2\right)\,+\\&& +\mbox{Li}_2\left(\frac{2w}{1+w}\right)-\mbox{Li}_2\left(-\frac{2w}{1-w}\right) -2\mbox{Li}_2\left(\frac{w}{1+w}\right)+2\mbox{Li}_2\left(-\frac{w}{1-w}\right) \,+\nonumber\\&& +\mbox{Li}_2\left(\frac{2aw}{b+aw}\right)-\mbox{Li}_2\left(-\frac{2aw}{b-aw}\right) -2\mbox{Li}_2\left(\frac{aw}{b+aw}\right)+2\mbox{Li}_2\left(-\frac{aw}{b-aw}\right) \nonumber\\ t_8&:=&\ln\left(\frac\xi 4\right)\ln\left(\frac{1+v}{1-v}\right) +\mbox{Li}_2\left(\frac{2v}{1+v}\right)-\mbox{Li}_2\left(-\frac{2v}{1-v}\right)-\pi^2\\ t_9&:=&2\ln\left(\frac{2(1-\xi)}{\sqrt\xi}\right)\ln\left(\frac{1+v}{1-v}\right) +2\left(\mbox{Li}_2\left(\frac{1+v}2\right)-\mbox{Li}_2\left(\frac{1-v}2\right)\right) \,+\nonumber\\&& +3\left(\mbox{Li}_2\left(-\frac{2v}{1-v}\right) -\mbox{Li}_2\left(\frac{2v}{1+v}\right)\right)\\ t_{10}&:=&\ln\left(\frac4\xi\right),\quad t_{11}\ :=\ \ln\left(\frac{4(1-\sqrt\xi)^2}\xi\right),\quad t_{12}\ :=\ \ln\left(\frac{4(1-\xi)}\xi\right) \end{eqnarray} \newpage
\section{Introduction} Laser interferometric detectors of gravitational waves such as the LIGO \cite{LIGO} and VIRGO \cite{VIRGO} are expected to be operational by the turn of the century. Gravitational waves from coalescing binary systems of black holes and neutron stars are relatively `clean' waveforms in the sense that they are easier to model and for this reason they are amongst the most important candidate sources for interferometric detectors. Binary systems are also valuable sources of astrophysical information as one can probe the universe up to cosmological distances. For instance, statistical analysis of several binary coalescences enables the estimation of the Hubble constant to an accuracy better than 10\% \cite {SCH86,Mark}. Events that produce high signal-to-noise ratio can be potentially used to observe such non-linear effects, such as gravitational wave tails, and to put general relativity into test in the strongly non-linear regime \cite {BS95}. Due to the weak coupling of gravitational radiation with matter the signal waveform is in general very weak and will not stand above the detector noise. In addition to the on-going efforts to reduce the noise, and hence increase the sensitivity of the detector, a considerable amount of research activity has gone into the development of efficient and robust data analysis techniques to extract signals buried in very noisy data. For a recent review on gravitational waves from compact objects and their detection see Thorne \cite{Th95a,Th95b}. Various data analysis schemes have been suggested for the detection of the `chirp' waveform from such systems \cite{MS95,Sm87,Th300}. Among them the technique of Weiner filtering is the most promising \cite{Th300,HEL,SCH89}. Briefly, this technique involves correlating the detector output with a set of templates, each of which is tuned to detect the signal with a particular set of parameters. This requires the signal to be known to a high level of accuracy. The fully general relativistic waveform from a coalescing binary system of stars is as yet unavailable. In the absence of such an exact solution, there have been efforts to find solutions perturbatively. Most of the work done in this area aims at computing the waveform correct to a high degree of accuracy so that theoretical templates computed based on them will obtain a large signal-to-noise ratio (SNR) when correlated with the detector output if the corresponding signal is present. In general, the number of parameters increases as we incorporate the higher order corrections. It is clear that the number of templates depends upon the number of signal parameters. As a consequence, the computing power for an on-line analysis will be greater for a larger number of parameters. In view of this restriction in computing power it is necessary to choose the templates in an optimal manner. This paper in part deals with this question. Investigations till now have been restricted to either choosing a finite subset of the signal space as templates \cite{SD91,DS94} or choosing templates from the `Newtonian' or the `first post-Newtonian' family of waveforms \cite{BD94,KKT94,A95,KKS95}. We generalize this problem by using the language of differential geometry. We show that it is unnecessary to restrict oneself to templates that are matched exactly to any particular signal. Differential geometry has been used in statistics before (see \cite{Am} and references therein) and the standard approach is to treat a set of parameterised probability distributions corresponding to a particular statistical model as a manifold. The parameters of the distribution serve as coordinates on this manifold. In statistical theory one frequently comes across the Fisher information matrix whose inverse gives a lower bound for the errors in the estimation of the parameters of a distribution. The Fisher information matrix turns out to be a very natural metric on the manifold of probability distributions and this metric can be used profitably in understanding the properties of a particular statistical model. Here in our paper we treat the set of coalescing binary signals corresponding to various parameters of the binary as a submanifold in the linear space of all detector outputs. We show in this paper that both the above mentioned manifolds are equivalent as far as their metrical properties are concerned. The geometric approach turns out to be useful not only in clarifying various aspects of signal analysis but also helps us to pose the question of optimal detection in a more general setting. Once a signal has been detected we can estimate the parameters of the binary. We assume that the parameters of the signal are the same as those of the template with which the maximum correlation is obtained. The errors involved in such an estimation has been worked out by several authors \cite{BS95,KKS95,F92,FC93,CF94,BS94,Kr,KLM93,PW95,JKKT}, for the case of `high' SNR and for the Newtonian and post-Newtonian waveforms using a single and a network of detectors. For the case of low SNRs one has to resort to numerical simulations. We have started a project to carry out exhaustive numerical simulations specifically designed to compute the errors in the estimation of parameters and covariances among them at various post-Newtonian orders, for circular and eccentric orbits, with and without spin effects and for different optical configurations of the interferometer. In this paper we report the results for the case of the initial LIGO configuration taking into account only the first post-Newtonian corrections and assuming circular orbits. Going beyond this needs tremendous amount of computing power which is just becoming available. The rest of the paper is organized as follows. In section \ref{waveforms} we describe the waveform from a coalescing binary system at various post-Newtonian orders. We introduce, following \cite{Sat94}, a set of parameters called `chirp times'. These parameters are found to be very convenient when we carry out Monte Carlo simulations. It turns out that the covariance matrix is independent of these parameters and hence it is sufficient to carry out the simulations only for a particular set of parameters. In section \ref{sm} we develop a geometric interpretation of signal analysis. We begin by introducing a metric on the manifold from a scalar product, which comes naturally from the theory of matched filtering and then show that this metric is the same as the one used by Amari \cite{Am}. Using the geometric approach we address the question of optimal filter placement and show that for the purpose of detection it is optimal to choose the templates outside the signal manifold. The covariance matrix of errors and covariances is shown to be the inverse of the metric on the manifold. In section \ref{sec_espar} we discuss the results of our simulations and compare the numerically obtained values and those suggested by the covariance matrix. We find substantial discrepancies in the predictions of the two methods. It is believed that the coalescing binary waveform shuts off abruptly at the onset of the plunge orbit. This has a major effect on the computations of the covariance matrix as well as on the Monte Carlo simulations. We discuss the effects of higher post-Newtonian corrections to the waveform. We also emphasisize the use of the instant of coalescence as a parameter in order to determine the direction to the source rather than the time of arrival \cite {BSD95}. Finally in section \ref{sec_con} we summarise our results and indicate future directions. \section{Coalescing Binary Waveforms} \label{waveforms} For the purpose of constructing templates for on-line detection, it is sufficient to work with the so called {\it restricted} post-Newtonian gravitational waveform. In this approximation the post-Newtonian corrections are incorporated only in the phase of the waveform, while ignoring corresponding corrections to the amplitude \cite {3mn}. Consequently, the restricted post-Newtonian waveforms only contain the dominant frequency equal to twice the orbital frequency of the binary computed up to the relevant order. In the restricted post-Newtonian approximation the gravitational waves from a binary system of stars, modeled as point masses orbiting about each other in a circular orbit, induce a strain $h(t)$ at the detector given by \begin {equation} h(t) = A (\pi f(t) )^{2/3} \cos \left [\varphi (t) \right ], \label {wave} \end {equation} where $f(t)$ is the instantaneous gravitational wave frequency, the constant $A$ involves the distance to the binary, its reduced and total mass, and the antenna pattern of the detector \cite{Th300}, and the phase of the waveform $\varphi (t)$ contains several pieces corresponding to different post-Newtonian contributions which can be schematically written as \begin {equation} \varphi(t) = \varphi_0(t) + \varphi_1(t) + \varphi_{1.5}(t) + \ldots. \label {phase} \end {equation} Here $\varphi_0(t)$ is the dominant Newtonian part of the phase and $\varphi_n$ represents the $n$th order post-Newtonian correction to it. In the quadrupole approximation we include only the Newtonian part of the phase given by \cite {Th300} \begin {equation} \varphi (t) = \varphi_0 (t) = {16 \pi f_a \tau_0 \over 5} \left [ 1 - \left ({f\over f_a}\right )^{-5/3} \right] + \Phi, \label {phaseN} \end {equation} where $f(t)$ is the instantaneous Newtonian gravitational wave frequency given implicitly by \begin {equation} t - t_a = \tau_0 \left [ 1 - \left ( {f \over f_a} \right )^{-8/3} \right ], \label {frequencyN} \end {equation} $\tau_0$ is a constant having dimensions of time given by \begin {equation} \tau_0 = {5 \over 256} {\cal M}^{-5/3} (\pi f_a)^{-8/3}, \label {NCT} \end {equation} and $f_a$ and $\Phi$ are the instantaneous gravitational wave frequency and the phase of the signal, respectively, at $t=t_a.$ The time elapsed starting from an epoch when the gravitational wave frequency is $f_a$ till the epoch when it becomes infinite will be referred to as the {\it chirp time} of the signal. In the quadrupole approximation $\tau_0$ is the chirp time. The Newtonian part of the phase is characterised by three parameters: (i) the {\it time of arrival} $t_a$ when the signal first becomes {\it visible} in the detector, (ii) the {\it phase} $\Phi$ of the signal at the time of arrival and (iii) the {\it chirp mass} ${\cal M} = (\mu^3 M^2)^{1/5},$ where $\mu$ and $M$ are the reduced and the total mass of the binary, respectively. At this level of approximation two coalescing binary signals of the same chirp mass but of different sets of individual masses would be degenerate and thus exhibit exactly the same time evolution. This degeneracy is removed when post-Newtonian corrections are included. When post-Newtonian corrections are included the parameter space of waveforms acquires an extra dimension. In this paper we show that even when post-Newtonian corrections up to relative order $c^{-4},$ where $c$ is the velocity of light, are included in the phase of the waveform it is possible to make a judicious choice of the parameters so that the parameter space essentially remains only three dimensional as far as the detection problem is concerned. It should, however, be noted that the evolution of the waveform must be known to a reasonably high degree of accuracy and that further off-line analysis would be necessary to extract useful astrophysical information. With the inclusion of corrections up to second post-Newtonian order the phase of the waveform becomes \cite {BDI} \begin {equation} \varphi (t) = \varphi_0 (t) + \varphi_1 (t) + \varphi_{1.5} (t) + \varphi_2 (t) \label {phasetotal} \end {equation} where $\varphi_0(t)$ is given by (\ref {phaseN}) and the various post-Newtonian contributions are given by \begin {equation} \varphi_1 (t) = 4 \pi f_a \tau_1 \left [ 1 - \left ( {f\over f_a} \right )^{-1} \right ] \label {phase1PN} \end {equation} \begin {equation} \varphi_{1.5} (t) = - 5 \pi f_a \tau_{1.5} \left [ 1 - \left ( {f\over f_a} \right )^{-2/3} \right ] \label {phase1.5PN} \end {equation} and \begin {equation} \varphi_2 (t) = 8 \pi f_a \tau_2 \left [ 1 - \left ( {f\over f_a} \right )^{-1/3} \right ]. \label {phase2PN} \end {equation} Now $f(t)$ is the instantaneous gravitational wave frequency correct up to second post-Newtonian order given implicitly by \begin {equation} t - t_a = \tau_0 \left [ 1 - \left ( {f \over f_a} \right )^{-8/3} \right ] + \tau_1 \left [ 1 - \left ( {f \over f_a} \right )^{-2} \right ] -\\ \tau_{1.5} \left [ 1 - \left ( {f \over f_a} \right )^{-5/3} \right ] + \tau_2 \left [ 1 - \left ( {f \over f_a} \right )^{-4/3} \right ]. \label {frequencyPN} \end {equation} In the above equations $\tau$'s are constants having dimensions of time which depend only on the two masses of the stars and the lower frequency cutoff of the detector $f_a$. The total chirp time now consists of four pieces: the Newtonian contribution $\tau_0$ is given by (\ref {NCT}) and the various post-Newtonian contributions are \begin {equation} \tau_1 = {5 \over 192\mu (\pi f_a)^2} \left ({743\over 336} + {11\over 4} \eta \right ), \end {equation} \begin {equation} \tau_{1.5} = {1\over 8 \mu} \left ( m \over \pi^2 f_a^5 \right)^{1/3} \end {equation} and \begin {equation} \tau_2 = {5\over 128 \mu} \left ({m\over \pi^2 f_a^2}\right )^{2/3} \left ({3058673 \over 1016064} + {5429\over 1008} \eta + {617\over 144}\eta^2\right) \end {equation} where $\eta=\mu/M.$ The phase (\ref {phasetotal}) contains the reduced mass $\mu$ in addition to the chirp mass $\cal M.$ Taking ($\cal M$, $\eta$) to be the post-Newtonian mass parameters the total mass and the reduced mass are given by $M = {\cal M} \eta^{-3/5},$ $\mu = {\cal M} \eta^{2/5}.$ Note that in the total chirp time $\tau$ of the signal the 1.5 post-Newtonian contribution appears with a negative sign thus shortening the epoch of coalescence. With the inclusion of higher order post-Newtonian corrections a chirp template is characterised by a set of four parameters which we shall collectively denote by $\lambda^\mu,$ $\mu=1,\ldots, 4.$ At the first post-Newtonian approximation instead of working with the parameters $\lambda^\mu =\{t_a,~\Phi,~{\cal M},~\eta\}$ we can equivalently employ the set $\{t_a,~\Phi,~\tau_0,~\tau_1\}$ for the purpose of constructing templates. This, as we shall see later, has some advantageous. However, at post-Newtonian orders beyond the first we do not have a unique set of chirp times to work with. The parameters $t_a$ and $\Phi$ are {\it kinematical} that fix the origin of the measurement of time and phase, respectively, while the Newtonian and the post-Newtonian chirp times are {\it dynamical} parameters in the sense that they dictate the evolution of the phase and the amplitude of the signal. It may be mentioned at this stage that in most of the literature on this subject authors use the set of parameters $\{t_C,~\Phi_C,~{\cal M},~ \eta\}$ where $t_C$ is the instant of coalescence and $\Phi_C$ is the phase of the signal at the instant of coalescence. In terms of the chirp times we have introduced, $t_C$ is the sum of the total chirp time and the time of arrival and $\Phi_C$ is a combination of the various chirp times and $\Phi:$ \begin {equation} t_C = t_a + \tau_0 + \tau_1 - \tau_{1.5} + \tau_2; \quad \end {equation} \begin {equation} \Phi_C = \Phi + {16\pi f_a \over 5} \tau_0 + 4 \pi f_a \tau_1 - 5\pi f_a \tau_{1.5} + 8\pi f_a \tau_2. \end {equation} In the stationary phase approximation the Fourier transform of the restricted first-post-Newtonian chirp waveform for positive frequencies is given by \cite {Th300,SD91,FC93,CF94} \begin {equation} \tilde h (f) = {\cal N} f^{-7/6} \exp \left [i\sum_{\mu=1}^6\psi_\mu(f)\lambda^\mu - i {\pi \over 4} \right ] \label {FT} \end {equation} where $${\cal N} = A\pi^{2/3} \left ( {2\tau_0 \over 3}\right )^{1/2} f_a^{4/3} $$ is a normalisation constant, $\lambda^\mu,$ $\mu=1,\ldots, 6,$ represent the various post-Newtonian parameters \begin {equation} \lambda^\mu = \left \{t_a, \Phi, \tau_0, \tau_1, \tau_{1.5}, \tau_2 \right \}, \end {equation} and \begin {eqnarray} \label {eqs1} \psi_1 & = & 2\pi f, \\ \psi_2 & = & -1, \\ \psi_3 & = & 2 \pi f -{ 16 \pi f_a \over 5}+ {6\pi f_a \over 5} \left ( {f\over f_a} \right )^{-5/3},\\ \psi_4 & = & 2\pi f - 4\pi f_a + 2\pi f_a \left ({f\over f_a}\right)^{-1},\\ \psi_5 & = &-2\pi f + 5\pi f_a - 3\pi f_a \left ({f\over f_a}\right)^{-2/3},\\ \psi_6 & = & 2\pi f - 8\pi f_a + 6\pi f_a \left ({f\over f_a}\right)^{-1/3}. \label {eqs2} \end {eqnarray} For $f<0$ the Fourier transform is computed using the identity $\tilde h(-f) = \tilde h^*(f)$ obeyed by real functions $h(t).$ In addition to the above mentioned parameters we shall introduce an amplitude parameter $\cal A$ in section \ref{sm}. \section{A geometric approach to signal analysis} \label{sm} In this section we apply the techniques of differential geometry to the problem of detecting weak signals embedded in noise. In section \ref{sms} we introduce the concept of the signal manifold and elaborate on the relationship of our approach with that of Amari \cite{Am}. Our discussion of the vector space of all detector outputs is modeled after the discussion given in \cite{SS94}. In section \ref{detect} we deal with the problem of choosing a set of filters for on-line analysis which would optimise the task of detection of the signal. In section \ref{2pnsec} we deal with the dimensionality of the chirp manifold when we incorporate higher order post-Newtonian corrections. It is found, that due to covariances between the parameters, it is possible to introduce an effective dimension which is less than the dimension of the manifold. This has very important implications for the detection problem. \subsection{Signal manifold} \label{sms} The output of a gravitational wave detector such as the LIGO, will comprise of data segments, each of duration $T$ seconds, uniformly sampled with a sampling interval of $\Delta$, giving the number of samples in a single data train to be $N = T/\Delta$. Each data train can be considered as a $N$-tuple $(x^0,x^1,\ldots,x^{N-1})$\, $x^i$ being the value of the output of the detector at time $i\Delta$. The set of all such $N$ tuples constitutes a $N$-dimensional vector space $\cal V$ where the addition of two vectors is accomplished by the addition of corresponding time samples. For later convenience we allow each sample to take complex values. A natural basis for this vector space is the {\em time basis} ${\bf e}_m^i = \delta^i_m$ where $m$ and $i$ are the vector and component indices respectively. Another basis which we shall use extensively is the Fourier basis which is related to the time basis by a unitary transformation $\hat U$: \begin{eqnarray} {\bf\tilde e}_m &=& \hat U^{mn}{\bf e}_n = \frac{1}{\sqrt N} \sum\limits_{n=0}^{N-1} {\bf e}_n\exp \left[\frac{2\pi i m n}{N}\right],\\ {\bf e}_m &=& \hat U^{\dagger mn}{\bf\tilde e}_n = \frac{1}{\sqrt N} \sum\limits_{n=0}^{N-1}{\bf\tilde e}_n\exp \left[-\frac{2\pi i m n}{N}\right]. \end{eqnarray} All vectors in $\cal V$ are shown in boldface, and the Fourier basis vectors and components of vectors in the Fourier basis are highlighted with a `tilde'. In the continuum case each data train can be expanded in a Fourier series and will contain a finite number of terms in the expansion, as the output will be band limited. The expansion is carried out over the exponential functions $\exp\left(2\pi i m t/T\right)$ which are precisely the Fourier basis vectors defined above. Though the index $m$ takes both positive and negative values corresponding to positive and negative frequencies it is both, possible and convenient to allow $m$ to take only positive values \cite{NUM}. Thus the vector space $\cal V$ can be considered as being spanned by the $N$ Fourier basis vectors implying immediately that the number of independent vectors in the time basis to be also $N$. This is the content of the Nyquist theorem which states that it is sufficient to sample the data at a frequency which is twice as large as the bandwidth of a real valued signal, where the bandwidth refers to the range of positive frequencies over which the signal spectrum is non zero. This factor of two does not appear in the vector space picture as we allow in general for complex values for the components in the time basis. A gravitational wave signal from a coalescing binary system can be characterised by a set of parameters $\bbox{\lambda} = (\lambda^0,\lambda^1,\ldots,\lambda^{p-1})$ belonging to some open set of the $p$-dimensional real space $R^p$. The set of such signals ${\bf s}(t;\bbox{\lambda})$ constitutes a $p$-dimensional manifold $\cal S$ which is embedded in the vector space $\cal V$. The parameters of the binary act as coordinates on the manifold. The basic problem of signal analysis is thus to determine whether the detector output vector $\bf x$ is the sum of a signal vector and a noise vector, ${\bf x} = {\bf s} + {\bf n}$, or just the noise vector, ${\bf x} = {\bf n}$, and furthermore to identify which particular signal vector, among all possible. One would also like to estimate the errors in such a measurement. In the absence of the signal the output will contain only noise drawn from a stochastic process which can be described by a probability distribution on the vector space $\cal V$. The covariance matrix of the noise $C^{ij}$ is defined as, \begin{equation} C^{ij} = \overline{n^in^{*j}}, \end{equation} where an * denotes complex conjugation and an overbar denotes an average over an ensemble. If the noise is assumed to be stationary and ergodic then there exists a noise correlation function $K(t)$ such that $C_{ij} = K(|i-j|\Delta)$. In the Fourier basis it can be shown that the components of the noise vector are statistically independent \cite{HEL} and the covariance matrix in the Fourier basis will contain only diagonal terms whose values will be strictly positive: $\tilde C_{ii} = \overline{\tilde n^i\tilde n^{*i}}$. This implies that the covariance matrix has strictly positive eigenvalues. The diagonal elements of this matrix $\tilde C_{ii}$ constitute the discrete representation of the power spectrum of the noise $S_n(f)$. We now discuss how the concept of matched filtering can be used to induce a metric on the signal manifold. The technique of matched filtering involves correlating the detector output with a bank of filters each of which is tuned to detect the gravitational wave from a binary system with a particular set of parameters. The output of the filter, with an impulse response $\bf q$, is given in the discrete case as \begin{equation} \label{corr} c_{(m)} = \frac{1}{\sqrt N}\sum\limits_{n=0}^{N-1}\tilde x^n \tilde q^{*n} \exp\left[-2\pi imn/N\right]. \end{equation} The SNR ($\rho$) at the output is defined to be the mean of $c_{(m)}$ divided by the square root of its variance: \begin{equation} \label{rho1} \rho \equiv \frac{\overline{c_{(m)}}}{\left[\overline{\left(c_{(m)} - \overline{c_{(m)}}\right)^2} \right]^{1/2}}. \end{equation} By maximising $\rho$ we can obtain the expression for the matched filter $\bf q_{(m)}$ matched to a particular signal ${\bf s}(t;\lambda^\mu)$ as \begin{equation} \label{matched} \tilde q^n_{(m)}(\lambda^\mu) = \frac{\tilde s^n(\lambda^\mu)\exp\left[2\pi imn/N\right]}{\tilde C_{nn}}, \end{equation} where $\rho$ has been maximised at the $m^{th}$ data point at the output and where $\mu = 1,2,\ldots,p$ where $p$ is the number of parameters of the signal. We now introduce a scalar product in $\cal V$. For any two vectors $\bf x$ and $\bf y$, \begin{equation} \label{scal} \left\langle{\bf x},{\bf s}\right\rangle = \sum\limits_{i,j=0}^{N-1}C^{-1}_{ij}x^iy^j = \sum\limits_{n=0}^{N-1}\frac{\tilde x^n \tilde s^{*n}(\lambda^\mu)}{\tilde C_{nn}}. \end{equation} In terms of this scalar product, the output of the matched filter $\bf q$, matched to a signal ${\bf s}(\lambda^\mu)$, can be written as, \begin{equation} \label{eqsc} c_{(m)}(\lambda^\mu) = \frac{1}{\sqrt N}\left\langle{\bf x},{\bf s}\right\rangle. \end{equation} As $\tilde C_{ii}$ is strictly positive the scalar product defined is positive definite. The scalar product defined above on the vector space $\cal V$ can be used to define a norm on $\cal V$ which in turn can be used to induce a metric on the manifold. The norm of a vector $\bf x$ is defined as $\|{\bf x}\| = \left\langle{\bf x},{\bf x} \right\rangle^{1/2}$. The ratio for the matched filter can be calculated to give $\rho = \left\langle{\bf s},{\bf s}\right\rangle^{1/2}$. The norm of the noise vector will be a random variable $\left\langle{\bf n},{\bf n}\right\rangle^{1/2}$ with a mean of $\sqrt{N}$ as can be seen by writing the expression for the norm of the noise vector and subsequently taking an ensemble average. The distance between two points infinitesimally separated on $\cal S$ can be expressed as a quadratic form in the differences in the values of the parameters at the two points: \begin{eqnarray} g_{\mu\nu}d\lambda^\mu d\lambda^\nu &\equiv& \|{\bf s}(\lambda^\mu + d\lambda^\mu) - {\bf s}(\lambda^\mu)\| = \|\frac{\partial{\bf s}}{\partial\lambda^\mu}d\lambda^\mu\|\\ &=& \left\langle\frac{\partial{\bf s}}{\partial\lambda^\mu}, \frac{\partial{\bf s}}{\partial\lambda^\nu}\right\rangle d\lambda^\mu d\lambda^\nu \end{eqnarray} The components of the metric in the coordinate basis are seen to be the scalar products of the coordinate basis vectors of the manifold. Since the number of correlations we can perform on-line is finite, we cannot have a filter corresponding to every signal. A single filter though matched to a particular signal will also detect signals in a small neighbourhood of that signal but with a loss in the SNR. The metric on the manifold quantifies the drop in the correlation in a neighbourhood of the signal chosen. Taking the output vector to be $\bf x$ and two signal vectors $\bf s(\bbox{\lambda})$ and $\bf s(\bbox{\lambda}+d\bbox{\lambda})$ and using Schwarz's inequality we have, \begin{eqnarray} \left\langle{\bf x},{\bf s(\bbox{\lambda}+d\bbox{\lambda})}\right\rangle - \left\langle{\bf x},{\bf s(\bbox{\lambda})}\right\rangle = \left\langle{\bf x},{\bf s}(\bbox{\lambda}+d\bbox{\lambda}) - {\bf s}(\bbox{\lambda}) \right\rangle &\le& \|{\bf x}\|\|{\bf s}(\bbox{\lambda} + d\bbox{\lambda}) - {\bf s}(\bbox{\lambda})\|\\ &=& \|{\bf x}\|g_{\mu\nu}d\lambda^\mu d\lambda^\nu. \end{eqnarray} As is apparent the drop in the correlation can be related to the metric distance on the manifold between the two signal vectors. We now discuss Amari's \cite{Am} work in the context of using differential geometry in statistics and elaborate on the relationship with the approach we have taken. The set of parametrised probability distributions corresponding to a statistical model constitute a manifold. The parameterized probability distributions in the context of signal analysis of gravitational waves from coalescing binaries are the ones which specify the probability that the output vector will lie in a certain region of the vector space $\cal V$ given that a signal ${\bf s}(t;\bbox{\lambda})$ exists in the output which we denote as $p({\bf x}| {\bf s}(t;\bbox{\lambda}))$ Since it is not our intention to develop Amari's approach any further we will be brief and will make all the mathematical assumptions such as infinite differentiability of functions, interchangeability of the differentiation and expectation value operators, etc. The set of probability distributions $p({\bf x}|{\bf s}(\bbox{\lambda}))$ where $\bbox{\lambda} \in R^p$, constitutes a manifold $\cal P$ of dimension $p$. At every point on this manifold we can construct a tangent space $T^0$ on which we can define the coordinate basis vectors as ${\bf\partial}_\mu = \frac{\bf\partial}{{\bf\partial}\lambda^\mu}$. Any vector $\bf A$ in this tangent space can be written as a linear combination of these coordinate basis vectors. We now define $p$ random variables $\sigma_\mu = -\frac{\partial}{\partial\lambda^\mu}\log(p({\bf x}|{\bf s}(\bbox{\lambda})))$. It can easily be shown that $\overline{\sigma_\mu} = 0$. We assume that these $p$ random variables are linearly independent. By taking all possible linear combinations of these random variables we can construct another linear space $T^1$. Each vector $\bf B$ in $T^1$ can be written as ${\bf B} = B^\mu\sigma_\mu$. The two vector spaces $T^0$ and $T^1$ are isomorphic to each other, which can be shown explicitly by making the correspondence $\sigma_\mu \leftrightarrow {\bf\partial}_\mu$. The vector space $T^1$ has a natural inner product defined on it which is the covariance matrix of the $p$ random variables $\sigma_\mu$. This scalar product can be carried over to $T^0$ using the correspondence stated above. The metric on the manifold can be defined by taking the scalar product of the coordinate basis vectors \begin{equation} \label{met} g_{\mu\nu} = \left\langle \partial_\mu,\partial_\nu\right\rangle = \overline{\sigma_\mu\sigma_\nu} \end{equation} In statistical theory the above matrix $g_{\mu\nu}$ is called the Fisher information matrix. We will also denote the Fisher matrix, as is conventional, as $\Gamma_{\mu\nu}$. It is clearly seen that orthogonality between vectors in the tangent space of the manifold is related to statistical independence of random variables in $T^1$. If we take the case of Gaussian noise the metric defined above is identical to the one obtained on the signal manifold by matched filtering. Gaussian noise can be described by the distribution, \begin{equation} \label{ndis} p({\bf n}) = \frac{\exp\left[ -\frac{1}{2}\sum\limits_{j,k=0}^{N-1}{C^{-1}_{jk} n^j n^{k*}}\right]} { \left[\ (2\pi)^N \det\left[C_{ij}\right]\ \right]^{1/2}} = \frac{\exp\left[ -\frac{1}{2}\sum\limits_{j,k=0}^{N-1}{\tilde C^{-1}_{jk} \tilde n^j \tilde n^{k*}}\right]} { \left[\ (2\pi)^N \det\left[\tilde C_{ij}\right]\ \right]^{1/2}} = \frac{\exp\left[ -\frac{1}{2}\sum\limits_{i=0}^{N-1}{\frac{\tilde n^i \tilde n^{i*}}{\tilde C_{ii}}}\right]} { \left[\ (2\pi)^N \det\left[\tilde C_{ij}\right]\ \right]^{1/2}}, \end{equation} where in the last step we have used the diagonal property of the matrix $\tilde C^{ij}$ which implies that $\tilde C^{-1}_{ii} = 1/\tilde C_{ii}$. As the noise is additive $p({\bf x}|{\bf s}(\bbox{\lambda}))$ can be written as $p({\bf x} - {\bf s}(\bbox{\lambda}))$. Assuming Gaussian noise we can write the expressions for the random variables $\sigma_i$ as, \begin{equation} \label{sig} \sigma_\mu = \left(\frac{1}{2}\right)\frac{\partial}{\partial\lambda^\mu}\left\langle{\bf x}-{\bf s}(\bbox{\lambda}), {\bf x} - {\bf s(\bbox{\lambda})}\right\rangle = \left\langle{\bf n},\frac{\partial}{\partial\lambda^\mu} {\bf s}(\bbox{\lambda})\right\rangle, \end{equation} where in the last step we have used ${\bf x} = {\bf s}(\bbox{\lambda})+{\bf n}$. The covariance matrix for the random variables $\sigma_\mu$ can be calculated to give \begin{equation} \label{sigco} \overline{\sigma_\mu\sigma_\nu} = \left\langle\frac{\partial}{\partial\lambda^\mu}{\bf s}(\bbox{\lambda}), \frac{\partial}{\partial\lambda^\nu}{\bf s}(\bbox{\lambda})\right\rangle, \end{equation} which is the same metric as defined over the signal manifold. Thus, both the manifolds $\cal S$ and $\cal P$ are identical with respect to their metrical properties. We will henceforth restrict our attention to the signal manifold $\cal S$. For the purpose of our analysis we will choose a minimal set of parameters characterizing the gravitational wave signal from a coalescing binary. We consider only the first post-Newtonian corrections. In section \ref{waveforms} we have already introduced the four parameters $\lambda^\mu = \{t_a,\Phi,\tau_0,\tau_1\}$. We now introduce an additional parameter for the amplitude and call it $\lambda^0 = \cal A$. The signal can now be written as $\tilde s(f;\bbox{\lambda}) = {\cal A} h(f;t_a,\Phi,\tau_0,\tau_1)$, where, $\bbox{\lambda} \equiv \{{\cal A},t_a,\Phi,\tau_0,\tau_1\}$. Numerically the value of the parameter $\cal A$ will be the same as that of the SNR obtained for the matched filter. We can decompose the signal manifold into a manifold containing normalised chirp waveforms and a one-dimensional manifold corresponding to the parameter $\cal A$. The normalised chirp manifold can therefore be parameterized by $\{t_a,\Phi,\tau_0,\tau_1\}$. This parameterization is useful as the coordinate basis vector $\frac{\partial}{\partial {\cal A}}$ will be orthogonal to all the other basis vectors as will be seen below. In order to compute the metric, and equivalently the Fisher information matrix, we use the continuum version of the scalar product as given in \cite{F92}, except that we use the two sided power spectral density. This has the advantage of showing clearly the range of integration in the frequency space though we get the same result using the discrete version of the scalar product. Using the definition of the scalar product we get \begin {equation} g_{\mu\nu} = \int_{f_a}^\infty {df\over S_n(f)} {\partial \tilde s(f; \bbox{\lambda})\over \partial \lambda^\mu} {\partial \tilde s^* (f; \bbox{\lambda}) \over \partial \lambda^\nu} + {\rm c.c.} \label{gamma2} \end {equation} Recall that in the stationary phase approximation the Fourier transform of the coalescing binary waveform is given by $\tilde h(f) = {\cal N} f^{-7/6} \exp \left [ i\sum_\mu\psi_\mu(f) \lambda^\mu \right]$ and $\tilde s(f) = {\cal A}\tilde h(f)$, where $\psi_\mu(f)$ are given by equations (\ref {eqs1}-\ref{eqs2}), $\mu=1,\ldots, 4$ and $\lambda^\mu = \{t_a,~\Phi,~\tau_0,~\tau_1\}.$ Note, in particular, that in the phase of the waveform the parameters occur linearly thus enabling a very concise expression for the components of $g_{\mu\nu}.$ The various partial derivatives are given by \begin {equation} {\partial \tilde s (f;\bbox{\lambda}) \over \partial {\lambda^\mu}} = i \psi_\mu (f) \tilde s(f;\bbox{\lambda}), \label{derivatives} \end {equation} where we have introduced $\psi_0= -i/{\cal A}.$ On substituting the above expressions for the partial derivatives in eq. (\ref {gamma2}) we get, \begin {equation} g_{\mu\nu} = \left < \psi_\mu h, \psi_\nu h \right > = 2 \int_{f_a}^\infty {{\psi_\mu(f) \psi^*_\nu(f) \left |\tilde h(f)\right|^2} \over S_n(f)} df \label{gamma3} \end {equation} The above definition of the amplitude parameter ${\cal A},$ as in Culter and Flannagan \cite {CF94}, disjoins the amplitude of the waveform from the rest of the parameters. Since $\psi_0$ is pure imaginary and $\psi_\mu$'s are real, it is straightforward to see from eq. (\ref {gamma3}) that \begin {equation} g_{00} = 1 \mbox{ and } g_{0\mu} = 0; \ \ \mu=1,\ldots,4. \end {equation} The rest of the components $g_{\mu\nu}$ are seen to be independent of all the parameters except $\cal A$ {\em i.e.} $g_{\mu\nu} \propto {\cal A}^2$. As ${\cal A}$ is unity for the normalised manifold the metric on the normalised manifold is flat. This implies not only that the manifold is intrinsically flat (in the stationary phase approximation) but also that the coordinate system used is Cartesian. If instead of the chirp time $\tau_0$ we use the parameter $\cal M$ then the metric coefficients will involve that parameter and the coordinate system will no longer remain Cartesian. \subsection{Choice of filters } \label{detect} We now use the formalism developed to tackle the issue of optimal filter placement. Till now, it has been thought necessary to use a finite subset of the set of chirp signals as templates for detection. We show that this is unduly restrictive. We suggest a procedure by which the detection process can be made more `efficient' by moving the filters out of the manifold. It must be emphasised that the algorithm presented below is both, simplistic and quite adhoc and is not necessarily the best. Moreover, we have implemented the algorithm only for the Newtonian case where the computational requirements are not very heavy. The signal manifold corresponding to higher post-Newtonian corrections will be a higher dimensional manifold and here the computational requirements will be substantial. The choice of optimal filters which span the manifold will then be crucial. Detection of the coalescing binary signal involves computing the scalar product of the output of the detector with the signal vectors. Subsequently one would have to maximise the correlations over the parameters and the number so obtained would serve as the statistic on the basis of which we can decide whether a signal is present in the given data train. Geometrically, this maximization corresponds to minimizing the angle between the output vector and the vectors corresponding to the normalised signal manifold. Using the cosine formula, \begin{equation} \cos(\theta) = \frac{\left<{\bf s}(\bbox{\lambda}),{\bf x}\right>}{\|{\bf x}\|\|{\bf s}(\bbox{\lambda})\|} = \frac{\|{\bf s}(\bbox{\lambda})\|^2 + \|{\bf x}\|^2 - \|{\bf x} - {\bf s}(\bbox{\lambda})\|^2} {2\|{\bf x}\|\|{\bf s}(\bbox{\lambda})\|}, \end{equation} it is clear that as $\|{\bf s}(\bbox{\lambda})\|$ is unity for the vectors belonging to the normalised signal manifold, maximising the scalar product is equivalent to minimising $ \|{\bf x} - {\bf s}(\bbox{\lambda})\|$ which is the distance between the tip of the output vector and the manifold. Given the constraints of computational power one would be able to evaluate only a finite number of these scalar products, say $n_F$, in a certain amount of time depending on the data train length and the padding factor. It is therefore necessary to be able to choose the $n_F$ filters in such a manner that the detection probability is maximal. We will need efficient on-line data analysis for two reasons: (i) To isolate those data trains which have a high probability of containing a signal and (ii) to determine the parameters of the binary early on during the inspiral and to use them for dynamical recycling techniques. Due to the finiteness of the filter spacing the signal parameters will in general not correspond to any of the $n_F$ filters chosen and this will lead to a drop in the maximum possible correlation. Till now attention has been focussed on identifying a set of optimal set of filters which are a discrete subset of the manifold. If detection is the sole purpose then the differential geometric picture suggests that confining the filter vectors to the signal manifold is an unnecessary restriction and in fact non optimal. Thus it is worthwhile to explore making a choice of filters outside the manifold. The filter vectors will thus belong to $\cal V$ but will not, in general, correspond to any signal. It is, of course, true that we are sacrificing on the maximum possible correlation obtainable (when the signal's parameters coincide with those of the filter). Thus the problem essentially is to select $n_F$ filter vectors which optimize the detection the efficiency of which depends upon the properties of the manifold. In general a single filter vector would have to pick up signals over a region of the manifold. The extent of this region is determined by fixing a threshold on the correlation between the filter and any signal in the region. We will denote this threshold by $\eta$, where $\eta$ takes a value which is close to, but less than unity. The typical value suggested for $\eta$ is $\sim 0.8$, \cite{SD91}. For a given filter $\bf q$ and a threshold $\eta$ the region on the manifold corresponding to the filter will be denoted as ${\cal S}_{\bf q}(\eta)$, where ${\cal S}_{\bf q}(\eta) \subset \cal S$. Geometrically this region is the intersection of an open ball in $\cal V$ of radius $2^{1/2}(1-\eta)^{1/2}$ (using the distance defined by the scalar product), with center $\bf q$, and the manifold. The $n_F$ filters taken together would have to `span' the manifold which means that the union of the regions covered by each filter would be the manifold itself {\em i.e.}, $\bigcup_{\bf q} {\cal S}_{\bf q}(\eta) = \cal S$. If the filter $\bf q$ lies on the manifold, then the correlation function $c_{\bf q}(\bbox{\lambda}) = \left<{\bf s}(\bbox{\lambda}), {\bf q}\right>$ will reach its maximum value of unity in ${\cal S}_{\bf q}(\eta)$ when ${\bf q} = {\bf s}(\bbox{\lambda})$ and will fall off in all directions. This means that the signals in the region which are further away from the filter are less likely to be picked up as compared to those in the immediate neighbourhood of the filter $\bf q$. We assume that a finite subset of the normalised signal manifold has been chosen to act as filters by some suitable algorithm \cite{SD91}, which taken together span the manifold. The number of filters will be determined by the available computing power. Consider one of these filters $\bf q$, the region corresponding to it for a threshold of $\eta$, ${\cal S}_{\bf q}(\eta)$, and an arbitrary normalised vector ${\bf q}_o$ which belongs to $\cal V$ but not necessarily $\cal S$. By correlating the vector ${\bf q}_o$ with vectors in ${\cal S}_{\bf q}(\eta)$ we obtain the correlation function $c_{{\bf q}_o}(\bbox{\lambda}) = \left<{\bf s}(\bbox{\lambda}),{\bf q}_o\right>$. We select that ${\bf q}_o$ to serve as a more optimal filter which maximises the average of this correlation function: \begin{equation} \left<{\bf s}(\bbox{\lambda}),{\bf q}_o\right>_{av} = \frac{1}{\int\limits_{{\cal S}_{\bf q}(\eta)} \sqrt{g}d^p\lambda}\int\limits_{{\cal S}_{\bf q}(\eta)}\left<{\bf s}(\bbox{\lambda}), {\bf q}_o\right>\sqrt{g}d^p\lambda= \left< \int\limits_{{\cal S}_{\bf q}(\eta)}{\bf s}(\bbox{\lambda})d^p\lambda\ ,\ {\bf q}_o\right>, \end{equation} where $g = det\left[g_{\mu\nu}\right]$. In the last step above the integration and the scalar product operations in the above equation have been interchanged. Moreover, for the normalised chirp manifold the metric does not depend upon the parameters in the coordinate system we have chosen and therefore, $g_{\mu\nu}$ is a constant and the factor $\sqrt{g}$ cancels. We now use Schwarz's inequality to maximise the average correlation to obtain, \begin{equation} \label{avcorr} {\bf q}_o = {\cal N}\int\limits_{{\cal S}_{\bf q}(\eta)}{\bf s}(\bbox{\lambda})d^p\lambda, \end{equation} where $\cal N$ is just a normalisation constant. We implemented the above algorithm for filter placement for the case of Newtonian signals with certain modifications. The normalised chirp waveform consists of three parameters $(\Phi,t_a,\tau_n)$. If we keep $t_a$ and $\tau_n$ fixed then the tip of the signal vector traces out a circle as we vary $\Phi$. As any circle lies on a plane we can express a signal vector as a linear sum of two vectors where the two vectors differ only in the phase parameter and we take this phase difference to be $\pi/2$. Thus we need only two mutually orthogonal filters to span the phase parameter. The time of arrival parameter $t_a$ is also a `convenient parameter' as by the use of fast Fourier transforms the correlations for arbitrary time of arrivals can be performed at one go. It is therefore not profitable for us to maximise the average correlation over the phase parameter $\Phi$ and the time of arrival $t_a$. In view of the above restrictions we modified the the filter placement algorithm. We consider the correlation function for the case when the filter vector is on the manifold. We define the `line of curvature' to be the curve on the manifold along which the correlation function falls the least. Figure (\ref{fig_dets}) illustrates the correlation function plotted as a function of $\tau_0$ along the line of curvature. It is seen from the contour diagram of the numerically computed correlation function that the line of curvature lies nearly on the submanifold $t_a + \tau_n = $ constant, of the normalised chirp manifold. We take two curves passing through the point $\bf q$ in the region ${\cal S}_{\bf q}(\eta)$: \begin{enumerate} \item $t_a + \tau_n = $ constant,\ \ $\Phi = 0$ and \item $t_a + \tau_n = $ constant,\ \ $\Phi = \pi/2$. \end{enumerate} We obtain one filter for each of the two curves by evaluating eq. \ref{avcorr} where the domains of integration correspond to the segments of the curves defined. Having determined the two filters we again plot the correlation function along the line of curvature as a function of $\tau_0$ in figure (\ref{fig_dets}). The region of the manifold selected corresponds to a range of $5.8$~secs to $6.0$~secs in the parameter $\tau_0$. It is interesting to note that we could have translated these values keeping the difference same without affecting our results. It can be seen that the correlation has a minimum at the center. In order to get a flatter correlation curve we select a linear combination of the original filter and the one obtained by averaging with suitable weights attached to each filter. This performs reasonably well as shown by the thick curve in the figure. The importance of having a flatter correlation function lies in the fact all the signals in a region can be picked up with equal efficiency and the drop in the maximum possible correlation can be compensated for by lowering the threshold. The average correlation obtained for the optimal filter is only marginally better than that obtained for the filter placed on the manifold. In the discussion above, we had started with a fixed number of filters $n_F$ on the manifold and obtained another set of $n_F$ filters which performs marginally better than the former set. Equivalently we can try to increase the span of each filter retaining the same threshold but reducing the number of filters required. In Figure \ref{fig_dets} we observe that the optimal filter chosen spans the entire region considered with a threshold greater than $0.9$, whereas the filter on the manifold spans about half the region at the same threshold. This indicates that by moving the filters out of the manifold in the above manner it may be possible to reduce the number of filters by a factor of two or so. One must however, bear in mind that the bank of filters obtained in this way are not optimal. There is scope to improve the scheme further. Our analysis is indicative of this feature. \subsection {Effective dimensionality of the parameter space of a second order post-Newtonian waveform} \label{2pnsec} It has already been shown that the first post-Newtonian waveform is essentially one-dimensional \cite{Sat94}. We argue in this subsection that even the second post-Newtonian waveform is essentially one-dimensional and a one-dimensional lattice suffices to filter the waveform. A Newtonian waveform is characterised by a set of three parameters consisting of the time of arrival, the phase of the signal at the time of arrival and chirp mass (equivalently, Newtonian chirp time). In this case, for the purpose of detection, one essentially needs to employ a one-dimensional lattice of filters corresponding to the chirp mass, the time of arrival being taken care by the fast Fourier transform algorithm and the phase being determined using a two-dimensional basis of orthogonal templates. When post-Newtonian corrections are included in the phase of the waveform the number of parameters increases apparently implying that one needs to use a two-dimensional lattice of filters corresponding to, say, the chirp and reduced masses (equivalently the Newtonian and post-Newtonian chirp times) which in turn means that the number of templates through which the detector output needs to be filtered goes up by several orders of magnitude. One of us (BSS) has recently shown that for the purpose of detection it is sufficient to use a one-dimensional lattice of filters even after first post-Newtonian corrections are included in the phase of the waveform and the relevant parameter here is the sum of the Newtonian and post-Newtonian chirp times. What happens when corrections beyond the first post-Newtonian order are incorporated in the phase of the waveform? The coalescing binary waveform is now available up to second post-Newtonian order \cite {BDI,BDIWW}. Blanchet et al. argue that the phase correction due to the second order post-Newtonian (2PN) term induces an accumulated difference of 10 cycles in a total of 16000. Consequently, it is important to incorporate the 2PN terms in the templates. When the 2PN terms are included it is useful to consider that the full waveform is parameterised by three additional parameters, corresponding to the chirp times at the 1, 1.5 and 2PN order (cf. Sec \ref{waveforms}), as compared to the Newtonian waveform. Of course, as far as the detection problem is concerned there is only one additional parameter since the chirp times are all functions of the two masses of the binary. However, for the purpose of testing general relativity one can consider each of the chirp times to be independent of the rest \cite{BS95,BS94}. Our problem now is to find the dimensionality of the parameter space of a 2PN waveform. To this end we consider the {\it ambiguity function} $C({\bbox {\lambda}}^{'},{\bbox {\lambda}})$ which is nothing but the correlation function of two normalised waveforms one of whose parameters (${\bbox {\lambda}}$) are varied by holding the parameters of the other fixed (${\bbox {\lambda}}^{'}$) : \begin {equation} C({\bbox {\lambda}}^{'}, {\bbox {\lambda}}) = \left < q({\bbox {\lambda}}^{'}), q({\bbox {\lambda}}) \right >; \ \ \left < q({\bbox {\lambda}}^{'}), q({\bbox {\lambda}}^{'}) \right > = \left < q({\bbox {\lambda}}^{\phantom{'}}),({\bbox {\lambda}}) \right > = 1. \end {equation} It is useful to think of ${\bbox {\lambda}}^{'}$ to be the parameters of a template and ${\bbox {\lambda}}$ to be that of a signal. With this interpretation the ambiguity function simply gives the span of a filter in the parameter space. The ambiguity function for the full waveform is a four-dimensional surface since there are four independent parameters. To explore the effective dimensionality of the parameter we consider the set of parameters to be $\{t_a, \Phi, m_1, m_2\},$ where $m_1$ and $m_2$ are the two masses of the binary. We have shown the contours of the ambiguity function maximised over $t_a$ and $\Phi$ (since these two parameters do not explicitly need a lattice of templates) in Fig. \ref{ambcont}. The template at the centre of the plot corresponds to a binary waveform with $m_1=m_2=1.4 M_\odot$ and the signal parameters are varied over the entire astrophysically interesting range of masses: $m_1, m_2 \in [1.4,10]M_\odot.$ From these figures we find that the ambiguity function is almost a constant along a particular line in the $m_1$--$m_2$ plane. This means that a template at the centre of the grid spans a relatively large area of the parameter space by obtaining a correlation very close to unity for all signals whose masses lie on the curve along which the ambiguity function roughly remains a constant. It turns out that the equation of this curve is given by \begin {equation} \tau_{0} + \tau_{1} - \tau_{1.5} + \tau_{2} = \mbox{const.} \end {equation} Let us suppose we begin with a two-dimensional lattice of filters corresponding to a certain grid (albeit, nonuniform) laid in the $m_1$--$m_2$ plane. Several templates of this set will have their total chirp time the same. Now with the aid of just one template out of all those that have the same chirp time we can effectively span the region that is collectively spanned by all such filters. More precisely, we will not have an appreciable loss in the SNR in replacing all templates of a given total chirp time by one of them. Consequently, the signal manifold can be spanned by a one-dimensional lattice of templates. \section {Estimation of parameters} \label{sec_espar} In this section we discuss the accuracy at which the various parameters of a coalescing binary system of stars can be estimated. All our results are for a single interferometer of the initial LIGO-type which has a lower frequency cutoff at 40 Hz. At present it is beyond the computer resources available to us to carry out a simulation for the advanced LIGO. In the first part of this section we briefly review the well known results obtained for the variances and covariances in the estimation of parameters using analytical methods. Analytical methods assume that the SNR is sufficiently large (the so-called strong signal approximation) and implicitly use a continuum of the parameter space. In reality, however, these assumptions are not necessarily valid and hence it is essential to substantiate the results obtained using analytical means by performing numerical simulations. In the second part of this section we present an exhaustive discussion of the Monte Carlo simulations we have performed to compute the errors and covariances of different parameters. As we shall discuss below the computation of errors using the covariance matrix is erroneous even at an SNR of 10-20. Our estimation of 1-$\sigma$ uncertainty in the various parameters, at low SNRs, is substantially larger than those computed using the covariance matrix. However, for high values of the SNR ($>25$--$30$) Monte Carlo estimation agrees with the analytical results. \subsection {Covariance matrix} \label {anacovar} In recent years a number of authors have addressed issues related to the variances expected in the parameter estimation \cite{BS95,KKS95,F92,FC93,CF94,BS94,Kr,KLM93,PW95,JKKT}, In the standard method of computing the variances in the estimation of parameters one makes the assumption that the SNR is so large that with the aid of such an approximation one can first construct the Fisher information matrix $\Gamma_{\mu\nu}$ and then take its inverse to obtain the covariance matrix $C_{\mu\nu}.$ In the strong signal approximation the Fisher information matrix and the covariance matrix are given by \begin {equation} g_{\mu\nu} = \Gamma_{\mu\nu} = \left < {\partial {\bf s} \over \partial \lambda^\mu}, {\partial {\bf s} \over \partial \lambda^\nu} \right >; \ \ \ C_{\mu\nu} = \Gamma^{-1}{\mu\nu}. \label{gamma1} \end {equation} As we have seen before the Fisher information and consequently the covariance matrix is block diagonal and hence there is no cross-talk, implying vanishing of the covariances between the amplitude and the other parameters. Consequently, we need not construct, for the purpose of Weiner filtering, templates corresponding to different amplitudes. The judicious choice of parameters has also allowed a very elegant expression for the Fisher information matrix. It is particularly interesting to note that the information matrix does not depend on the values of the various parameters, except for the amplitude parameter, and hence is a constant as far as the parameter space of the normalised waveforms is considered. For the purpose of numerical simulations it is convenient to choose the set $\lambda^\mu = \{{\cal A}, t_a, \Phi, \tau_0, \tau_1\}$ where $\cal A$ is the amplitude parameter, $t_a$ and $\Phi$ are the time of arrival of the signal and its phase at the time of arrival, and $\tau_0$ and $\tau_1$ are the Newtonian and the post-Newtonian coalescence times. For noise in realistic detectors, such as LIGO, the elements of the Fisher information matrix cannot be expressed in a closed form and, for the set of parameters employed, it is not useful to explicitly write down the covariance matrix in terms of the various integrals since the errors and covariances do not have any dependence on the parameters. We thus evaluate the information matrix numerically and then take its inverse to obtain the covariance matrix. Instead of dealing with the covariance matrix $C$ it more instructive to work with the matrix of standard deviations and correlation coefficients $D$ which is related to the former by \begin {equation} {D_{\mu\nu}} = \left \{ \begin{array} {ll} \sqrt {C_{\mu\nu}}, & {\rm if} \ \mu=\nu\\ C_{\mu\nu}/(\sigma_\mu \sigma_\nu) & {\rm if}\ \mu \ne \nu, \end {array}\right. \label {varcorr} \end {equation} where $\sigma_\mu={D}_{\mu\mu}$ is the 1-$\sigma$ uncertainty in the parameter $\lambda_\mu.$ The off-diagonal elements of $D$ take on values in the range $[-1, 1]$ indicating how two different parameters are correlated: For $\mu\ne\nu,$ ${D}_{\mu\nu}=1$, indicates that the two are perfectly correlated, ${D}_{\mu\nu}=-1$ means that they are perfectly anticorrelated and ${D}_{\mu\nu}=0$ implies that they are uncorrelated. Since the information matrix is block-diagonal, the amplitude parameter is totally uncorrelated with the rest and thus an error in the measurement of $\cal A$ will not reflect itself as an error in the estimation of the other parameters and vice versa. In contrast, as we shall see below, Newtonian chirp time is strongly anticorrelated to post-Newtonian chirp time, which implies that if in a given experiment $\tau_0$ happens to be estimated larger than its true value then it is more likely that $\tau_1$ will be estimated to be lower than its actual value. Such correlations are useful as far as detection is concerned since they tend to reduce the number of templates needed in filtering a given signal. On the other hand, strong correlations increase the volume in the parameter space to which an event can be associated at a given confidence level. It seems to be in general true that a given set of parameters do not satisfy the twin properties of having small covariances and reducing the effective dimension of the manifold for the purpose of filtering. We elaborate on this point below. Given a region in a parameter space, it is useful to know the proper volume (as defined by the metric) of the manifold corresponding to the said region. In choosing a discrete set of filters for the detection problem one has to decide upon the maximum allowable drop in the correlation due to the finite spacing. Once this is fixed, the number of filters can be determined from the total volume of the manifold. For the detection problem it is beneficial to have a small volume whereas if the waveform is parameterized in such a way such that the manifold corresponding to it covers a large volume, then one can determine the parameters to a greater accuracy. As a simple example let us consider a two-dimensional toy model: $\bbox{\lambda} = \{\lambda_1,\lambda_2\}$. We compare different signal manifolds each corresponding to a different parameterizations of the waveform. We assume that the covariance matrix and its inverse, the Fisher information matrix, to be: \begin{equation} \label{volumes} C_{\mu\nu} = \left( \begin{tabular}{cc} $\sigma_{11}$&$\sigma_{12}$\\ $\sigma_{12}$&$\sigma_{22}$ \end{tabular} \right) \mbox{\ \ \ and\ \ \ } \Gamma_{\mu\nu} = \left( \begin{tabular}{cc} $\gamma_{11}$&$\gamma_{12}$\\ $\gamma_{12}$&$\gamma_{22}$ \end{tabular} \right) = \frac{1}{\left(\sigma_{11}\sigma_{22} -\sigma_{12}^2\right)} \left( \begin{tabular}{cc} $\sigma_{22}$&$-\sigma_{12}$\\ $-\sigma_{12}$&$\sigma_{11}$ \end{tabular} \right). \end{equation} The volume of the manifold corresponding to a region $\cal K$ of the parameter space is given as, \begin{equation} \label{vol2} V_{\cal K} = \int_{\cal K}\gamma_{11}\gamma_{22} \left[1 - \frac{\gamma_{12}^2}{\gamma_{11}\gamma_{22}}\right]d\lambda_1d\lambda_2 = \int_{\cal K}\gamma_{11}\gamma_{22}\left[1 - \epsilon\right]d\lambda_1d\lambda_2, \end{equation} where $\epsilon=\gamma_{12}^2/(\gamma_{11} \gamma_{22})$ is the correlation coefficient. It can be clearly seen that if the correlation coefficient is small, keeping the variances in the parameters, fixed then the volume of the manifold is maximal. Since the parameters $\tau_0$ and $\tau_1$ are highly anticorrelated the proper volume corresponding to the region reduces to zero showing that the effective dimensionality of the manifold is less. Though, in principle, the variances and covariances are independent of the chirp time, in reality there arises an indirect dependence since one terminates a template at a frequency $f=1/(6^{3/2}\pi M)$ (where $M$ is the total mass of the binary) corresponding to the plunge radius at $a=6M.$ Therefore, larger mass binaries are tracked over a smaller bandwidth so much so that there is less frequency band to distinguish between two chirps of large, but different, total mass. Consequently, at a given SNR the error in the estimation of chirp times is larger for greater mass binaries. This is reflected by the fact that the integrals in eq. (\ref {gamma2}) are somewhat sensitive to the value of the upper cutoff. (This also explains why the errors in the estimation of the chirp and reduced masses are larger for greater mass binaries \cite{FC93,CF94}.) In the following we assume that the noise power spectral density is that corresponding to the initial LIGO for which a fit has been provided by Finn and Chernoff \cite{FC93}. For an SNR of 10 the matrix $D$ is given by \begin {equation} {D_{\mu\nu}} = \left ( \begin{array} {rrrrr} 1.0 & 0 & 0 & 0 \\ & 8.37 & 0.999 & -0.999 \\ & & 3.16 & -0.998 \\ & & & 8.4 \\ \end {array} \right ), \end {equation} for the Newtonian signal, and by \begin {equation} {D_{\mu\nu}} = \left ( \begin{array} {rrrrr} 1.0 & 0 & 0 & 0 & 0 \\ & 20.4 & 0.997 & -0.972 & 0.911 \\ & & 6.7 & -0.954 & 0.881 \\ & & & 45.1 & -0.982 \\ & & & & 25.98 \\ \end {array} \right ), \end {equation} for the first post-Newtonian corrected signal. While computing varianaces and covariances the integrals in equation (\ref {gamma3}) are evaluated by chosing a finite upper limit of 1 kHz. In the above matrices the entries are arranged in the order $\{{\cal A}, t_a, \Phi, \tau_0\}$ in the Newtonian case, $\{{\cal A}, t_a, \Phi, \tau_0, \tau_1\}$ in the post-Newtonian case, off diagonal elements are dimensionless correlation coefficients and, where appropriate, diagonal elements are in ms. The values quoted in the case of the Newtonian waveform are consistent with those obtained using a different set of parameters by Finn and Chernoff \cite {FC93}. In Fig.~\ref {fdependence} we have plotted $\sigma$'s, at an SNR of 10, as a function of the upper frequency cutoff $f_c$ for Newtonian and post-Newtonian chirp times and the instant of coalescence $t_C.$ We see that $\sigma$ is larger for higher mass binaries but this is because we have fixed the SNR. However, if we consider binaries of different total masses, all located at the same distance, then a more massive binary produces a stronger SNR so that in reality it may be possible to determine its parameters more accurately than that of a lighter binary. In Fig.~\ref{truedependence} we have plotted $\sigma$'s for binaries all located at the same distance as a function of total mass. We fix one of the masses at a value of $1.4~M_\odot$ and vary the other from $1.4~M_\odot$ to $10~M_\odot$. In computing the $\sigma$'s plotted in this figure we have terminated the waveform at the plunge orbit and normalised the SNR of a $10 M_\odot$-$1.4 M_\odot$ binary system to 10. As a function of $M$ the uncertainties in $\tau_0$ and $\tau_1$ initially falloff since the increase in the SNR for larger mass binaries more than compensates for the drop in the upper frequency cutoff. However, for $M$ larger than a certain $M_0$ the increase in SNR is not good enough to compensate for the drop in $f_c$ so much so that the uncertainties in $\tau_0$ and $\tau_1$ increase beyond $M_0$. The parameter $t_a$, however, falls off monotonically. \subsection {Monte Carlo estimation of parameters} In this Section we present the first in a series of efforts to compute the covariance matrix of errors through numerical simulations for a coalescing binary waveform at various post-Newtonian orders. Analytical computation of the covariance matrix, as in the previous section, gives us an idea of the covariances and variances but, as we shall see in this Section, at low SNR's it grossly underestimates the errors. Quite apart from the fact that the assumptions made in deriving the covariance matrix might be invalid at low SNR's, in a realistic detection and data analysis, other problems, such as discreteness of the lattice of templates, finite sampling of the data, etc., do occur. It therefore seems necessary to check the analytical calculations using numerical simulations to gain further insight into the accuracy at which physical parameters can be measured. This Section is divided into several parts: In the first part we highlight different aspects of the simulation, in the second part we briefly discuss the choice of templates for the simulation, in the third we elaborate on the Monte Carlo method that we have adopted to carry out our simulations and in the fourth we discuss problems that arise in a numerical simulation. The results of our study are discussed in the next Section. \subsubsection {Parameters of the simulation} Let $s(t)$ be a signal of strength $\cal A$ characterised by a set of parameters $\hat{\bbox{\lambda}}$ \begin {equation} s(t; \hat{\bbox{\lambda}}) = {\cal A} h(t;\hat{{\bbox {\lambda}}}),\ \ \left<h,h\right >=1. \end {equation} In data analysis problems one considers a discrete version $\{s^k | k=0,\ldots,N-1\}$ of the waveform $s(t)$ sampled at uniform intervals in $t:$ \begin {equation} s^k \equiv s(k\Delta); \ \ \ k=0,\ldots,N-1, \end {equation} where $\Delta$ denotes the constant interval between consecutive samples and $N$ is the total number of samples. The sampled output $x^k$ of the detector consists of the samples of the noise plus the signal: \begin {equation} x^k = n^k + s^k. \end {equation} The {\it sampling rate,} $f_s=\Delta^{-1}$ (also referred to as the {\it sampling frequency}) is the number of samples per unit time interval. In a data analysis problem the sampling frequency is determined by the signal bandwidth. If $B$ is the signal bandwidth, i.e., if the Fourier transform of the signal is only nonzero over a certain interval $B,$ then it is sufficient to sample at a rate $f_s=2B.$ In our case there is a lower limit in the frequency response of the detector since the detector noise gets very large below a seismic cutoff at about 10--40 Hz. As mentioned in the last Section there is also an upper limit in frequency up to which a chirp signal is tracked since one does not accurately know the waveform beyond the last stable orbit of the binary. This corresponds to gravitational wave frequency $f_c=1/(6^{3/2}\pi M).$ For a neutron star-neutron star (ns-ns) binary $f_c \sim 1525$~Hz while for a ns-black hole (of 10 $M_\odot$) (ns-bh) binary $f_c \sim 375$~Hz. Due to constraints arising out of limited computational power, we terminate waveforms at $750$ Hz even when $f_c$ is larger than 1000 Hz. Such a shutoff is not expected to cause any spurious results since, even in the case of least massive binaries of ns-ns, which we consider in this study, more than 99 \% of the `energy' is extracted by the time the signal reaches 750 Hz. We have carried out simulations with two types of upper cutoff: \begin {enumerate} \item one in which all templates, irrespective of their total mass, are shutoff beyond 750 Hz., \item a second in which the upper frequency cutoff is chosen to be $750$ Hz or $f_c,$ whichever is lower. \end {enumerate} Consistent with these cutoffs the sampling rate is always taken to be $2$ kHz. (We have carried out simulations with higher sampling rates and found no particular advantage in doing so nor did we find appreciable changes in our results.) In all our simulations, as in the previous section, we take the detector noise power spectral density $S$ to be that corresponding to initial LIGO \cite {FC93}. For the purpose of simulations we need to generate noise corresponding to such a power spectrum. This is achieved by the following three steps: \begin {enumerate} \item generate Gaussian white noise $n'^k$ with zero mean and unit variance, $$\overline {n'^k}=0,\ \ \overline{n'^k n'^l} = \delta^{kl}$$ where an overbar denotes average over an ensemble, \item compute its Fourier transform $$\tilde n'^k\equiv \frac{1}{\sqrt{N}}\sum_{l=0}^{N-1} n'^l \exp (2\pi i k l / N)$$ and \item multiply the Fourier components by the square root of the power spectral density, $$\tilde n^k = \sqrt {S^k} \tilde n'^k.$$ \end {enumerate} The resultant random process has the requisite power spectrum. In the above, the second step can be eliminated since the Fourier transform of a Gaussian random process is again a Gaussian, but with a different variance. In other words we generate the noise directly in the Fourier domain. The simulated detector output, in the presence of a signal $s^k,$ in the Fourier domain is given by \begin {equation} \tilde x^k = \tilde n^k + \tilde s^k \end {equation} where $\tilde s^k$ is the discrete Fourier transform of the signal. \subsubsection {Choice of templates} To filter a Newtonian signal we employ the set of parameters $\{t_a, \Phi, \tau_0\}$ and to filter a post-Newtonian signal we employ the set $\{t_a, \Phi, \tau_0, \tau_1\}.$ Templates need not explicitly be constructed for the time of arrival since computation of the scalar product in the Fourier domain (and the availability of fast Fourier transform (FFT) algorithms) takes care of the time of arrival in essentially one computation ($N\log_2 N$ operations as opposed to $N^2$ operations, where $N$ is the number of data points). Moreover, there exists a two-dimensional basis for the phase parameter which allows the computation of the best correlation with the aid of just two filters. Consequently, the parameter space is essentially one-dimensional in the case of Newtonian signals and two-dimensional in the case of post-Newtonian signals. (However, as shown in Section \ref {2pnsec} it is to be noted that for the purpose of detection the effective dimensionality of the parameter space, even with the inclusion of second post-Newtonian corrections, is only one-dimensional.) We adopt the method described in Sathyaprakash and Dhurandhar \cite {SD91} to determine the templates needed for chirp times. As described in \cite {SD91,Sat94} filters uniformly spaced in $\tau_0$ and $\tau_1$ covers the parameters space efficiently. \subsubsection {Monte Carlo method} \label {mcmethod} In order to compute variances and covariances numerically, we employ the Monte Carlo method. The basic idea here is to mimic detection and estimation on a computer by performing a very large number of simulations so as to minimize the uncertainties induced by noise fluctuations. In our simulations we generate a number of detector outputs $\{x^k\}$ each corresponding to a definite signal $s^k(\hat{{\bbox {\lambda}}})$ of a certain strength, but corresponding to different realisations of the random process $\{n^k\}.$ Computation of the covariance matrix involves filtering each of these detector outputs through an a priori chosen set (or lattice) of templates $\{q(t; {{\rm _t}\bbox{\lambda}_k}) | k=1,\ldots,n_f\},$ where $n_f$ denotes the number of templates. The templates of the lattice each has a distinct set of values of the {\em test} parameters ${{\rm _t}\bbox{\lambda}_k}$ and together they span a sufficiently large volume in the parameter space. The simulated detector output is correlated with each member of the lattice to obtain the corresponding filtered output $C ({{\rm _t}\bbox{\lambda}}_k):$ \begin {equation} C({{\rm _t}\bbox{\lambda}}_k) = \left < x, q ({{\rm _t}\bbox{\lambda}}_k) \right >. \end {equation} For a given realisation of noise a particular template obtains the largest correlation and its parameters are the {\it measured} values ${{\rm _m}{\bbox{\lambda}}}$ of the signal parameters. Thus, the measured values of the parameters are defined by \begin {equation} \max_k C({{\rm _t}\bbox{\lambda}}_k) = C({{\rm _m}{\bbox{\lambda}}}). \end {equation} The measured values, being specific to a particular realisation of noise, are random variables. Their average provides an estimation ${{\rm _e}{\bbox{\lambda}}}$ of the true parameter values and their variance is a measure of the error $\sigma_{\bbox {\lambda}}$ in the estimation: \begin {equation} {{\rm _e}{\bbox {\lambda}}} = \overline{{\rm _m}{\bbox {\lambda}}},\ \ \ \ \ \sigma_{\bbox {\lambda}}^2 = \overline{ \left ( { {{\rm _m}{\bbox {\lambda}}} } - \overline{ {{\rm _m}{\bbox {\lambda}}}} \right )^2}, \ \ \ \ D^{\mu\nu} = \overline{ {{\rm _m}\lambda}^\mu \ {{\rm _m}\lambda}^\nu \over \sigma_\mu \sigma_\nu}, \ \ (\mu\ne \nu), \label {mccovar1} \end {equation} where $D_{\mu\nu}$ are the correlation coefficients between parameters $\lambda_\mu$ and $\lambda_\nu.$ In order to accurately determine $\sigma_{\bbox {\lambda}}$ a large number of simulations would be needed. If the measured values ${{\rm _m}{\bbox {\lambda}}}$ obey Gaussian statistics then after $N_{\rm t}$ trials the variance is determined to a relative accuracy $1/\sqrt N_{\rm t}$ and estimated values can differ from their true values by $\sigma_{\bbox{\lambda}}/\sqrt N_{\rm t}.$ We have performed in excess of 5000 trials, for each input signal, and thus our results are accurate to better than 1 part in 70. Even more crucial than the number of simulations is the number of templates used and their range in the parameter space. We discuss these and other related issues next. The actual templates chosen, say for the parameter $\tau_0,$ in a given `experiment' depend on the true parameters of the signal, the number of noise realisations employed and the expected value of the error. Let us suppose we have a first guess of the error in $\tau_0$, say $\sigma_{\tau_0}$. Then, we choose 51 uniformly spaced filters around $\hat {\tau}_0$ (where $\hat{\tau_0}$ is the signal chirp time) such that: \begin{equation} {{\rm _t}\tau_0} \in \left [ \hat{\tau}_0 - 5 \sigma_{\tau_0}, \hat{\tau}_0 + 5 \sigma_{\tau_0} \right ]. \end{equation} This implies that we are covering a 5$\sigma$ width in $\tau_0$ at a resolution of $\sigma_{\tau}/5.$ The probability, that a template between 4$\sigma$ and 5$\sigma$ from the true signal `clicks', being $\sim 6\times 10^{-5},$ we are on safe grounds since, in a given simulation, we consider no more than 5000 trials. (In comparison, the probability that a template between 3$\sigma$ and 4$\sigma$ clicks is $2.2\times 10^{-3}$ corresponding to an expected 13 events in 5000 trials.) For a post-Newtonian signal, which in effect needs to be spanned by a two-dimensional lattice of filters, the above choice of templates implies a requirement of $2601 \times 2$ filters in all. Here a factor of 2 arises because for each filter in the $\tau_0$--$\tau_1$ space we will need two templates corresponding to the two independent values of the phase $\Phi$: $0$ and $\pi/2.$ In the case of a Newtonian signal, the lattice being one-dimensional, one can afford a much higher resolution and range. Even with the aid of a mere 201 templates we can probe at a $\sigma/10$ resolution with a $10\sigma$ range. We start off a simulation with the pretension that there is no knowledge of what the $\sigma_{{\bbox {\lambda}}}$'s are. Thus, we choose as our first trial a very large $\sigma_{\bbox {\lambda}}$ and lay the lattice of templates. With this lattice we perform a test run of 400 trials and examine the distribution of the measured values. If the distribution is not, as expected, a Gaussian then we alter $\sigma_{\bbox {\lambda}}$: we decrease it if the distribution is too narrow and increase it if the distribution is too wide and does not show the expected falloff. In particular, we make sure that the templates at the boundary of the chosen range do not click even once and the skewness of the distribution is negligible. When for a certain $\sigma_{\bbox {\lambda}}$ a rough Gaussian distribution is observed then we carry out a simulation with a larger number of trials (typically 5000). We subject the measured values in this larger simulation to the very same tests described above. We only consider for further analysis such simulations which `pass' the above tests and determine the estimates, variances and covariances of the parameters using the measured values, with the aid of equation (\ref {mccovar1}). \subsubsection {Numerical errors and remedies} There are several sources of numerical errors that tend to bias the results of a simulation unless proper care is exercised to rectify them. In this Section we point out the most important ones and show how they can be taken care of. Due to memory restrictions, the present version of our codes work with single precision except the FFT, which is implemented in double precision. In future implementations we plan to carry out all computations in double precision. This will possibly reduce some of the numerical noise that occurs, especially at high SNRs, in the present simulations. \begin {enumerate} \item {\it Orthonormality of filters:} For the sake of simplicity it is essential that the filters are normalised in the sense that their scalar product is equal to unity: $ \left < q, q \right > = 1. $ A waveform is normalized numerically using the discrete version of the scalar product: \begin {equation} {\cal N} = {1 \over \sum_{k=0}^{N-1} S^{-1}_k \left | \tilde q^k \right |^2 } \end {equation} As mentioned earlier we use a two-dimensional basis of filters for the phase parameter. Choosing the two filters to be orthogonal to each other makes the maximisation over phase easier. However, here care must be exercised. Two filters $q(t; t_a, \tau_0, \tau_1, \Phi=0)$ and $q(t; t_a, \tau_0, \tau_1, \Phi=\pi/2)$ are apparently orthogonal to each other. The numerically computed `angle' between the two filters, chosen in this manner, often turns out to be greater than $\sim 10^{-2}$ radians. Consequently, one obtains erroneous correlations. In order to circumvent this problem we first generate two filters that are roughly orthogonal to each other, as above, and then use the Gram-Schmidt method to orthogonalise the two vectors. If an explicit numerical orthogonalisation such as this is {\em not} implemented then the measured values of the various parameters show spurious oscillations in their distribution and the estimated values of the parameters tend to get biased. \item {\it Correlation function:} The scalar product of two normalised templates $q(t; {{\rm _t}{\bbox {\lambda}}})$ and $q(t; {{\rm _t}{\bbox {\lambda}}}')$ is given by \begin {equation} C({{\rm _t}\lambda}_\mu, {{\rm _t}\lambda}_\nu') = \left < q(t; {{\rm _t}\lambda}_\mu), q(t; {{\rm _t}\lambda}_\nu') \right >, \ \ \left < q(t; {{\rm _t}{\bbox {\lambda}}}), q(t; {{\rm _t}{\bbox {\lambda}}}) \right > = \ \ \left < q(t; {{\rm _t}{\bbox {\lambda}}}'), q(t; {{\rm _t}{\bbox {\lambda}}}') \right > = 1, \end {equation} where we have indicated the dependence of the scalar product on the various parameters by explicitly writing down the parameter subscripts. Let us fix the parameters of one of the templates, say ${{\rm _t}{\bbox {\lambda}}},$ and vary the parameters of the other template. Of particular interest is the behaviour of $C$ maximised over all but one of the parameters, say $\lambda_{\nu_0}$: \def\ooo #1 #2{\vphantom{S}^{\raise-0.5pt\hbox{$#1$}}_{\raise0.5pt \hbox{$#2$}}} \begin {equation} C_{\rm max}({ {\rm _t}\lambda}_\mu, {{\rm _t} \lambda}_{\nu_0}') = \max_{\ooo {\scriptstyle \ {{\rm _t}\lambda}_\nu'} {\scriptstyle\nu \ne \nu_0}} C ({{\rm _t}\lambda}_\mu, {{\rm _t} \lambda}_\nu'). \end {equation} $C_{\rm max}$ is expected to drop monotonically as $|\lambda_{\nu_0} - \lambda_{\nu_0}'|$ increases. However, we have observed departures from such a behaviour possibly arising out of numerical noise. Such a behaviour causes bias in the estimation of parameters, and consequently in the determination of their covariances, especially at high SNRs. We have found no remedy to this problem and some of our results at high SNRs may have biases introduced by this effect. (Sampling the templates at a higher rate did not help in curing this problem.) \item {\it Grid effects:} The parameters of a signal chosen for the purpose of simulation and detection can in principle be anything and in particular it need not correspond to any of the templates of the lattice. However, in practice we find that whenever the signal parameters do not correspond to a member of the lattice then the resultant simulation has a bimodal distribution of the measured values. This is, of course, expected since a signal not on the grid is picked up by two nearest templates along each direction in the parameter space. Sometimes we do find that the peaks corresponding to the bimodal distribution does not belong to the nearest neighbour filters but slightly away. This is related to the fact that the correlation function maximised over the time of arrival and the phase of the signal falls off much too slowly along the $\tau_0$--$\tau_1$ direction and small deviation from a monotonic fall can cause biases. (Such biases would be present in the case when a signal corresponds to one of the grid points though the magnitude of the effect would be lower.) In order to avoid this problem, and the consequent shifts in the estimation of parameters and errors in the determination of variances and covariances, we always choose the parameters of the signal to be that corresponding to a template. \item {\it Upper frequency cutoff and its effect on parameter estimation} The Fisher information matrix computed using the stationary phase approximation in Section \ref{sm} does not include the effect of truncating the waveform at $a = 6M$ --- the plunge cutoff. As mentioned before, we have carried out simulations for both with, and without, incorporating the upper cutoff. As the covariance matrix incorporating the upper cutoff is not available we have been able to compare the Monte Carlo results with the covariance matrix only for the latter case, where the cutoff is held fixed at $750$~Hz. If we incorporate the upper cutoff into the Monte Carlo simulations the errors in the parameters are reduced drastically. The effect of the upper cutoff is expected to be more important for the higher mass binaries such as the ones we have considered. The ambiguity function, in this case, no longer remains independent of the point on the manifold. In other words, the correlation between two chirps depends not only on the difference between the parameters of the signals, but also on the absolute values of the parameters. The correlation surface also ceases to be symmetric {\em i.e.} the correlation between two chirps also depends on the sign of $\delta\lambda$, where, $\delta\lambda$ is the difference in the values of the parameters. As the computational power required for carrying out simulations for lower mass binaries is not available to us the simulations have been restricted to ns-bh star binaries, where the effect of the upper cutoff is important. \item {\it Boundary effects} For the purpose of simulations a grid of filters has to be set up `around' the signal. The grid must be large enough so that the estimated parameters do not overshoot the boundary of the grid. This causes a problem as every value in the \{$\tau_0,\tau_1$\} plane does not lead to a meaningful value for the masses of the binary system. This does not however prevent us to construct a waveform with such a value for \{$\tau_0,\tau_1$\} even though the signal in general does not correspond to any `real' binary system. This is valid, and even necessary, if we are to compare the numerical results with the covariance matrix. \item {\it Incorporating the cutoff in the presence of boundary effects} If we wish to incorporate the effects of the upper cutoff in simulations then we run into a serious problem, as we would have to know the total mass of the binary in order to compute the upper cutoff. For an arbitrary \{$\tau_0,\tau_1$\} we can end up with negative and even complex values of the total mass and hence the upper cutoff at $a = 6M$ is not meaningful. Thus, we cannot even construct a waveform for an arbitrary combination of \{$\tau_0,\tau_1$\}. Therefore, in such cases, we restrict ourselves to simulations where the grid lies entirely within valid limits for \{$\tau_0,\tau_1$\}. \end {enumerate} \subsection {Results and Discussion} Our primary objective is to measure the variances and covariances following the method described in Sec. \ref{mcmethod} and study their departure from that predicted by analytical means (cf. Sec. \ref {anacovar}). We have carried out simulations for several values of the masses of the binary and in each case the signal strength (which is a measure of the SNR) is varied in the range 10--40. However, since the variances and covariances are independent of the absolute values of the parameters, for the parameter set that we employ, results are only quoted corresponding to a typical binary system. (See Sec. \ref {anacovar} for a discussion of the covariance matrix.) Similar results are obtained in other cases too. We use two sets of parameters to describe our results. Monte Carlo simulations allow us to directly measure the amplitude, the time of arrival, the phase at the time of arrival and the chirp time(s). This is the set $\{{\cal A}, t_a, \Phi, \tau_0, \tau_1\}.$ As we shall see below, the instant of coalescence can be measured much more accurately than the time of arrival. As a consequence of this, the direction to the source can be determined at a {\em much greater accuracy} by employing $t_C$ as a parameter instead of $t_a.$ Thus, we also quote estimates and errors for the parameter $t_C.$ Since the error in the estimation of the phase is quite large, even at high SNRs, we ignore it in our discussions. We first deal with the Newtonian signal and highlight different aspects of the simulation and discuss the results at length. We then consider the first post-Newtonian corrected signal. \subsubsection {Newtonian signal} In the case of Newtonian signals the parameter space is effectively one-dimensional and, as mentioned earlier, in this case the lattice of templates covers a $10$--$\sigma$ range of the parameters at a resolution of $\sigma/10$ centred around the true parameters of the signal. In Fig. \ref {nmcfig} we have shown the error $\sigma_{\bbox {\lambda}}$ in the estimation of parameters $t_a,$ $\tau_0$ and $t_C,$ as a function of SNR, deduced using the covariance matrix as solid lines and computed using Monte Carlo simulations as dotted lines. The curve corresponding to the covariance matrix is obtained using an upper frequency cutoff $f_c= 750$~Hz consistent with that used in our simulations. The errorbars in the estimation of $\sigma_{{\bbox {\lambda}}}$'s are obtained using 4 simulations, each with 4000 trials. At low SNR's $\sigma_{{\bbox {\lambda}}}$'s have a larger uncertainty, as expected, and for $\rho > 30$ this uncertainty is negligible, and sometimes smaller than the thickness of the curves, except in the case of $\sigma_{t_C}$ (see below, for a possible explanation). At low SNRs (10--15) there is a large departure of the various $\sigma_{\bbox {\lambda}}$'s from that inferred using the covariance matrix. At an SNR $\sim 17$ the two curves merge (except in the case of $\sigma_{t_C}$) indicating the validity of the covariance matrix results for this and higher SNR's. Interestingly, the agreement between Monte Carlo simulation results and those obtained using the covariance matrix is reached roughly at the same SNR irrespective of the parameter in question. We note that in spite of the fact that the time of arrival and the chirp time have large errors, the instant coalescence can be estimated very accurately---an order of magnitude better than either. What is puzzling, however, is that, in the case of $t_C,$ the Monte Carlo curve drops below the covariance matrix curve above an SNR of 15 and the two curves do not seem to converge to one another even at very high SNR's. Coincident with the crossover of the two curves, the error in the estimation of $\sigma_{t_C}$ increases, contrary to what happens for the other parameters, signalling that there is a large fluctuation in the estimation of $\sigma_{t_C}.$ This behaviour, we guess, is an artifact of the low value of the sampling rate. Of course, our sampling rate is sufficiently high to respect the sampling theorem. However, since $t_C$ is determined to an accuracy an order of magnitude better than either $t_a$ or $\tau_0,$ a much higher resolution in template-spacing would be needed for determining the error in the instant of coalescence than that used for estimating the errors in the time of arrival or the chirp time(s). Testing this claim, unfortunately, is beyond the computer resources at our disposal since we would need a sampling rate of about 10 kHz with a filter-spacing $10^{-4}$~s. We hope to be able to resolve this issue in course of time. Nevertheless, the fact that the error in the estimation of $\sigma_{t_C}$ first decreases with SNR and increases only after the two curves crossover, hints at the above possibility as a cause for this anomalous behaviour. This effect is also observed in the case of a post-Newtonian signal. In Table \ref{nmctab} we have given the actual signal parameters $\hat {{\bbox {\lambda}}},$ estimated values of the parameters ${{\rm _e}{\bbox {\lambda}}}$ (cf. equation (\ref{mccovar1})) and the corresponding errors in their estimation $\sigma_{\bbox {\lambda}},$ for several values of the SNR. Errors inferred from the covariance matrix can be read off from Fig.~\ref{nmcfig}. The estimated values are different from the true values, some of them being overestimated and some others being underestimated. However, the deviations are often larger than what we expect. In a simulation that uses $N_{\rm t}$ trials the estimated parameters ${{\rm _e}{\bbox {\lambda}}},$ assuming a Gaussian distribution for the measured parameters ${{\rm _m}{\bbox {\lambda}}},$ can be different from the true values by $\sigma_{\bbox {\lambda}}/\sqrt{N_{\rm t}}.$ (In contrast, the measured values ${{\rm _m}{\bbox {\lambda}}}$ can differ from their true values by $\sigma_{\bbox {\lambda}}$ or more.) However, we often obtain a slightly larger deviation, \begin {equation} 2 {\sigma_{\bbox {\lambda}}\over \sqrt {N_{\rm t}}} \la \left |{{\rm_e}{\bbox {\lambda}}} - \hat{{\bbox {\lambda}}}\right | \la 3 {\sigma_{\bbox {\lambda}}\over \sqrt {N_{\rm t}}}, \end {equation} and we are unable to resolve this discrepancy. A more concrete test for the simulations is the histogram $n({{\rm _m}{\bbox {\lambda}}})$ of the measured parameters, namely the frequency at which a given parameter clicks in a simulation. This is shown plotted in Fig.~\ref {nhist} for an SNR of 10. The skewness of the measured value is less than $10^{-2}.$ These results lend further support to the Monte Carlo simulations. There are visible asymmetries in the distributions of $\tau_0$ and $t_a$ and the asymmetries in the two cases are of opposite sense. This can, of course, be understood from the fact that $t_a$ and $\tau_0$ have a negative correlation coefficient. The histogram of $t_C,$ even at an SNR of 10, has very few non-zero bins. This of course reflects the fact that it is determined very accurately. We are unable to resolve the central peak in $n(t_C)$ since, as mentioned earlier, the sampling rate and resolution in $\tau_0$ are not good enough to do so. \subsubsection {Post-Newtonian signal} As opposed to the Newtonian case here we have essentially a two-dimensional lattice of filters corresponding to $\tau_0$ and $\tau_1.$ For the purpose estimating variances and covariances we lay a mesh consisting of $2601 \times 2$ uniformly spaced filters around the true parameters of the signal. As pointed out in Sec. \ref {detect} not all filters in the mesh, unlike in the Newtonian case, would correspond to the waveform from a realistic binary but that does not preclude their use in the Monte Carlo simulations. We shall see that the results of our simulations lend further support to the claim that for the purpose of detection, the parameter space need only be one-dimensional \cite {Sat94}. The results obtained for the first post-Newtonian signal are qualitatively similar to that of a Newtonian signal and we refer the reader, where appropriate, to the Newtonian case for a more complete discussion. In Fig. \ref {pnmcfig} we have shown the error in the estimation of parameters $\tau_0,$ $\tau_1,$ $t_C$ and $t_a,$ clockwise from top left, respectively, as a function of SNR. The solid and dotted curves are as in Fig. \ref{nmcfig}. Here again the upper frequency cutoff is taken to be 750 kHz. Just as in the case of a Newtonian signal here too the results obtained from Monte Carlo simulation are much higher than those obtained by employing the covariance matrix. At an SNR of 10, which is the expected value for a majority of events that initial LIGO and VIRGO interferometers will observe, the Monte Carlo values are more than thrice as much as their corresponding covariance matrix values and at an SNR of 15 the errors are roughly twice that expected from the covariance matrix. In absolute terms, however, the errors are still quite small compared to the actual parameter values: for a ns-ns binary, at an SNR of 10, \begin {equation} {\sigma_{\tau_0}\over \tau_0} \sim 2.4 \%, \ \ {\sigma_{\tau_1}\over \tau_1} \sim 9.4 \%. \end {equation} At an SNR of 10 the time of arrival can be measured to an accuracy of 72 ms in contrast to a value of 20 ms expected from the covariance matrix. As is well known, with the inclusion of the post-Newtonian terms error in the estimation of the time of arrival and Newtonian chirp time increases by about a factor of 2 and 3, respectively \cite{FC93,CF94}. As in the Newtonian case here again we see that the Monte Carlo curves approach the corresponding covariance matrix curves at a high SNR the only difference being that the agreement is reached at a higher SNR $\sim 25.$ For SNRs larger than this the two curves are in perfect agreement with each other. As mentioned earlier, $\sigma_{t_C}$ shows an anomalous behaviour possibly arising out of insufficient resolution in the time of arrival and the chirp times. In Table \ref {pntab} we have listed the true parameters $\hat{{\bbox {\lambda}}},$ the estimated values ${{\rm _e}{\bbox {\lambda}}}$ and the Monte Carlo errors $\sigma_{\bbox {\lambda}}$ for different SNRs. As in the Newtonian case here too the estimated values show a larger departure, than expected, from the true values. Histograms of the various measured parameters including $t_C$ are shown in Fig. \ref {pnhist} for a signal strength of 10. The skewness is below its standard deviation of $\sqrt {15/N_{\rm t}}$ \cite {NUM} indicating the Gaussian nature of the various distributions. Even in the case of a post-Newtonian signal $\sigma_{t_C}$ is so small that we only have three non-zero bins in $n(t_C).$ We now turn attention to other, more general, issues arising out of the simulations. In Sec. \ref {2pnsec} we have argued, on the basis of the behaviour of the noise-free correlation function, that the effective dimensionality of the parameter space for the purpose of detection, even in the case of a post-Newtonian signal, is only one-dimensional. The results of our Monte Carlo simulation unambiguously show that this is indeed true even in the presence of noise. We investigated the two-dimensional histogram, that gives the number of occurrances of different templates in the lattice, in a particular simulation. The templates that `click' are all aligned along the line $\tau_0 + \tau_1=$~const. In a total of 5000 realisations corresponding to this simulation there is only one instance when a filter outside this region clicks giving a probability of less than $10^{-3}$ for a template outside this region to give a maximum. Consequently, it is only necessary to choose a single filter along this curve. The distribution of the maximum correlation $ C_{\rm max}(\hat {{\bbox{\lambda}}}, { {\rm _t}{\bbox {\lambda}}})$ obtained from different noise realisations needs a special mention since it has an inherent bias. In Fig. \ref {snrhist} we have shown the distribution of the maximum correlation taken from one of our simulations corresponding to an SNR of 10. Notice a slight shift of the distribution towards a higher value and this cannot be accommodated within the expected fluctuation in the mean. The measured value of the standard deviation $\sigma_{\cal A}$ is $0.95.$ Since the number of simulations is 5000 we expect that the signal strength should differ from the true value of 10 by no more than $\sigma_{\cal A}/\sqrt {5000}=0.014.$ However, the mean value is 10.26 giving a deviation of $0.26$ which is about 20 times larger than that expected. This occurs at all SNRs and for both Newtonian and post-Newtonian signals. This of course does not mean there is bug in the way we are computing the maximum correlation. In the process of maximisation, values greater than the signal strength are favoured and consequently the mean of the maximum correlation shows a shift towards a higher value. This suggests that that the maximum of the correlation is a biased estimator of the signal strength. We find, consistent with the covariance matrix calculation, that the amplitude parameter is uncorrelated with the rest of the parameters; crosscorrealtion coefficients $D_0\mu,$ $\mu\ne 0,$ (cf. eq. \ref {varcorr}) inferred from our Monte Carlo simulations are less than $\sim 10^{-6}.$ Finally, it is of interest to note how the phase parameter $\Phi$ is correlated with the time of arrival. A plot of ${{\rm _m}\Phi}$ versus ${{\rm _m}t}_a$ is shown in Fig. \ref{taphi}. We find that the measured values of the time of arrival and the phase are such that $2\pi f_{0} \ {{\rm _m}t}_a = {{\rm _m}\Phi},$ where $f_{0}$ has a value of approximately $51$~Hz. When the time of arrival shifts by more than a cycle of the signal the phase jumps by a factor $2 \pi$ leading to the points seen in the top-left and bottom-right corner of the figure. This makes the estimation of the phase and the error in its estimation pretty involved. \subsubsection {Incorporating the effects of upper cutoff} As mentioned before, incorporating an upper cutoff at the onset of the plunge has a drastic effect on the estimation of parameters. The incorporation of the upper cutoff is implemented by stopping the waveform when the instantaneous frequency reaches the frequency associated with the onset of the plunge or $750$~Hz, whichever is lower. However, due to computational constraints we have carried out the simulations only for high mass binaries and hence the upper cutoff plays an important role in all our simulations. It is to be noted that the discussion of the ambiguity function in section \ref{2pnsec} is not valid when the effect of the upper cutoff is taken into account though for low mass binaries, such as ns-ns binaries, the results there will still be valid. The further dependence of the signal waveform on the total mass of the system through the upper cutoff means that we can estimate the individual masses more exactly though the computational power is bound to increase. In order to carry out the simulations for the present case we selected a $10 M_\odot-1.2 M_\odot$ binary system as this enables us to choose the filter grid well within valid limits of $\tau_0$ and $\tau_1$. The simulations were carried out for various values of SNR starting from 10. The histograms of the estimated parameters at an SNR of 10 is shown in fig. \ref{fig_phistu}. At this SNR the errors obtained are $\sigma_{\tau_0} = 39.3$ ms, $\sigma_{\tau_1} = 22.4$ ms, $\sigma_{t_a} = 23.1$ ms and $\sigma_{t_C} = 0.6$ ms. These can be compared with the values in table \ref{pntab} and we can see that except for the parameter $t_C$ the errors are substantially lesser when the upper cutoff is incorporated into the waveform. It is necessary to recompute the covariance matrix, as emphasized before, including the effect of the upper cutoff in order to compare these numerically obtained values with the covariance matrix. In order to do this it is not enough to replace the upper limit in the integral in eq. (\ref{gamma3}) with the upper cutoff. The waveform now depends on the total mass of the system through the upper cutoff and this information has to be incorporated into the waveform. We carried out the simulations for various SNRs for the same value of masses quoted above. In the absence of the estimates of the covariance matrix when the upper cutoff is incorporated we assume that at an SNR of 40 the Monte Carlo estimates are consistent with those of the covariance matrix. In Figure~\ref{fig_psigu} we illustrate the results of our simulations. The points on the dotted line are the values obtained through Monte Carlo estimates and the continuous line is obtained by fitting a $1/\rho$ dependence of the errors on the SNR {\em assuming } consistency at an SNR of 40. It is seen as in the previous simulations that except for the parameter $t_C$ the other parameters are fairly consistent with a $1/\rho$ dependence of the errors on the SNR at an SNR greater than 25. \section{Conclusions} \label{sec_con} In this paper we have introduced the use of differential geometry in studying signal analysis and have addressed issues pertaining to optimal detection strategies of the chirp waveform. We have also carried out Monte Carlo simulations to check how well the covariance matrix estimates the errors in the parameters of the chirp waveform. We summarize below our main results. \begin{enumerate} \item We have developed the concept of a signal manifold as a subset of a finite dimensional vector space of detector outputs. Using the correlation between two signal vectors as a scalar product we have induced a metric on the signal manifold. With this geometric picture it is possible to pose the question of optimal detection in a more general setting. We suggest that the set of template waveforms for the detection of the chirp signal need not correspond to any point on the chirp manifold. We propose an algorithm to choose templates off the signal manifold and show that the drop in the correlation due to the discreteness of the set of templates is reduced. This algorithm, though certainly not the best, motivates the search of more efficient templates. In addition, the chirp manifold corresponding to the second post-Newtonian waveform is shown to be effectively one-dimensional. This has important implications for the computational requirement for the on-line detection of the chirp signal. The use of a convenient set of parameters of the chirp waveform for carrying out numerical and analytical simulations is stressed. These parameters are such that the metric components are independent of the parameters which implies that the manifold is flat and the corresponding `coordinate system' is Cartesian. As the metric defined is nothing but the Fisher information matrix, the covariance matrix, being the inverse of the Fisher information matrix, is also independent of the parameters. \item Monte Carlo simulations have been carried out for the case of the initial LIGO to find out whether the actual errors in the estimation of parameters is consistent with the values predicted by the covariance matrix. Simulations have been carried out for both the Newtonian as well as the post-Newtonian waveforms. We have restricted ourselves to the case of high mass binary systems, such as bh-ns binaries, where the computational requirement is not very heavy since the length of the data train, in such cases, works out to be less than $8$~s. Nevertheless, as has been shown in this paper, the covariance matrix is independent of the parameters identified by us when waveforms are terminated at a constant upper cutoff irrespective of their masses. Consequently, our results will hold good for binary systems of arbitrary masses. We point out the major problems that arise while performing a numerical simulation and, where appropriate, we suggest how they may be taken care of. In particular, the effect of incorporating the upper cutoff in the frequency of the gravitational wave at the onset of the plunge, which essentially depends on the total mass of the binary, is extremely important for high mass binaries. Since the covariance matrix with the inclusion of such a mass-dependent upper cutoff is not available we have carried out most of our simulations using a constant upper cutoff. This enables us to directly compare the results of our Monte Carlo simulations with those of the analytically computed covariance matrix. Since for binaries with total mass less than $5 M_\odot$ the plunge induced upper cutoff is larger than that induced by the detector noise these effects can be ignored for such binaries. The numerical experiments indicate that the covariance matrix underestimates the errors in the determination of the parameters even at SNRs as high as 20. In the Newtonian case the correlation coefficient of the time of arrival $t_a$ and the Newtonian chirp time $\tau_0$ is found to be very close to $-1,$ so much so that even at a SNR of $7.5$, the instant of coalescence $t_C = t_a + \tau_0$ remains practically a constant. The error in the estimation of $\tau_0$ for the post-Newtonian waveform is about four times the error obtained in the case of the Newtonian waveform at the same SNR. This is expected as the first post-Newtonian correction to the waveform introduces a new parameter $\tau_1$ (called the first post-Newtonian chirp time) which is highly (anti)correlated with $\tau_0$. For the post-Newtonian waveform at an SNR of $10$ the error in $\tau_0$ is about $3$ times that predicted by the covariance matrix. This corresponds to a factor of $2$ in the chirp mass $\cal M.$ The distributions for the parameters have been obtained and are seen to be unimodal distributions and are slightly more sharper than a Gaussian. When the plunge induced upper cutoff is incorporated into the waveform the errors in the estimation of parameters decrease by a factor of about $2.5$. The correlation coefficient between $\tau_0$ and $\tau_1$ is also found to decrease, which is consistent with our discussion in Section \ref{anacovar}. The results obtained suggest that higher moments than that used in obtained in computing the covariance matrix may be important in the determination of the errors in parameter estimation. In the geometric picture this amounts to taking into account curvature effects, either intrinsic or extrinsic. \item We suggest that $t_C$ is a more suitable parameter to estimate the direction to the source than the time of arrival. The latter is a kinematical parameter that fixes the time at which the gravitational wave frequency reaches the lower cutoff of the detector while the parameter $t_C$ has the physical significance of being the instant of coalescence. At an SNR of 10 the error in $t_a$ is too large (20 ms) to deduce the direction to the source accurately, whereas the error in the parameter $t_C$ is less than $0.5$~ms. This will further go down substantially for the advanced LIGO. A detailed analysis of the determination of the direction to the binary using delays in $t_C$ between different detectors is carried out in Bhawal and Dhurandhar \cite{BD95} (also see \cite {BSD95}). \end{enumerate} We now suggest further work which needs to be done along the lines of this paper. A full understanding of the chirp signal manifold when higher post-Newtonian corrections are incorporated into the waveform is in order. This will help in the development of more efficient algorithms for the choice of templates in the detection problem and facilitate reduction in computational time. The Monte Carlo simulations which we have carried our are for the case of a binary waveform correct up to first post-Newtonian order. Moreover, only circular orbits are considered. The effect of eccentricity is currently being investigated \cite{GBS95}. Performing simulations when higher post-Newtonian corrections are taken into account calls for an immense amount of computational time. Fortunately, matched filtering algorithm being amenable to parallelization \cite {SD93}, one could aim at using the massively parallel computers, that are now becoming available the world over, in performing such simulations. \acknowledgments The authors would like to thank the members of the gravitational wave group at IUCAA especially S.D. Mohanty and B. Bhawal for many helpful discussions. R.B. is being supported by the Senior Research Fellowship of CSIR, India.
\section*{\Large\bit 1) Introduction} \par Although the Kerr-geometry, when considered classically, represents a vacuum solution of the Einstein-equations, there have been various attempts \cite{Israel1,Israel2,Burinskii,Lopez} to find a singular matter distribution as its source. The natural candidate for locating such an energy-momentum distribution is the singular region of the spacetime.\par Using distributional techniques it is indeed possible to find a tensor-distribution \cite{BaNa2}, which is supported in the singular region and which from the distributional viewpoint represents the right-hand side of the Einstein equations. A central ingredient in this endeavour is the Kerr-Schild decomposition of the metric which is the main reason for the applicability of distributional techniques to the non linear Einstein equations. Moreover, the flat (background) part of the decomposition provides us with a natural notion of boosts as being its associated isometries. However, as already noted in the Schwarzschild-case \cite{AiSe}, boosting the metric itself does not produce a sensible result. Therefore the strategy is to shift from the metric to the energy-momentum tensor which has a well-defined limit \cite{BaNa1}. The resulting limit is a pp-wave \cite{JEK}. Solving the Einstein-equations with this inhomogeneity produces the so-called Aichelburg--Sexl (AS) geometry describing the gravitational field of a massless point-particle. The result is independent of the direction of the boost which is due to the spherical symmetry of the original geometry. \par This is, however, no longer the case for Kerr, which is only axis-symmetric. In a first step \cite{BaNa3} the authors considered therefore a boost along the preferred direction, namely the axis of symmetry. The aim of the present work is to lift this restriction and to investigate the form of the limit by boosting along an arbitrary direction. Finally we will discuss the extremal case where the boost direction becomes perpendicular to the axis of symmetry in some detail. \section*{\Large\bit 2) The general boost} The main ingredient of our approach is the Kerr-Schild decomposition of the Kerr geometry \begin{equation}\label{KerrSchild} g_{ab}=\eta_{ab} + f \>k_a k_b, \end{equation} where $\eta_{ab}$ denotes the flat background part, with respect to which boosts find their natural home. $k^a$ denotes a geodetic null vector field and $f$ a scalar function. With respect to Kerr-Schild coordinates \cite{HaEl} $\eta_{ab}$ becomes manifestly flat and $k^a$ and $f$ are given by ($\rho^2 =x^2 + y^2 $) \begin{eqnarray*} k^a=(1,k^i), && k^i = \frac{r\rho}{r^2+a^2} e^i_\rho - \frac{a\rho}{r^2+a^2} e^i_\phi \\ f=\frac{2 m r}{\Sigma} , && \Sigma = \frac{r^4 + a^2 z^2}{r^2} \end{eqnarray*} where $r$ is subject to $r^4-r^2 (\rho^2 +z^2-a^2) -a^2 z^2 =0.$ The Ricci-tensor for geometries in the Kerr-Schild class takes the form \begin{equation}\label{Ricci} R^a{}_b = \frac{1}{2} (\partial^a\partial_c(fk^c k_b) + \partial_b\partial_c(fk^ck^a) - \partial^2(fk^a k_b) ), \end{equation} which in the Kerr case gives rise to the distributional energy-momentum tensor \cite{BaNa2} \begin{align}\label{KerrEM} T^a{}_b &= \frac{m\delta(z)}{8\pi}\left\{ \frac{2}{a}\left( \left[ \frac{a^2\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} ^3}\right ] -\frac{\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} } - \delta(\rho- a) \right )(dt)^a(\partial_t)_b \right. \nonumber\\ &+((\partial_t)^a(e_\phi )_b - (e_\phi )^a (dt)_b)\left( 2\left [\frac{\rho\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} ^3}\right ] -\frac{\pi}{a}\delta(\rho- a) \right ) \nonumber\\ &+\frac{2}{a}\left( - \left[\frac{\rho^2\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} ^3}\right ] - \frac{\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} } + 2\delta(\rho- a) \right )(e_\phi )^a(e_\phi )_b \nonumber\\ &\left. -\frac{2}{a}\frac{\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} }(e_\rho )^a(e_\rho )_b\right \} \end{align} The square-bracket terms in (\ref{KerrEM}) represent distributional extensions of the corresponding non-locally integrable functions to the whole of test function space. Their definition may be exemplified by $$ \left(\left[ \frac{\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} ^3} \right] ,\varphi \right) := \int\limits_{\rho\leq a} d^2x\frac{1}{\sqrt{a^2-\rho^2} ^3 }(\varphi(x) -\varphi(a e_\rho)), $$ where $e_\rho$ denotes the radial unit-vector with respect to polar coordinates. For a more detailed discussion on the origin of these terms the reader is referred to \cite{BaNa2}. Interpreting Kerr-Schild coordinates as being asymptotically at rest we may rewrite $T^a{}_b$ with respect to an arbitrary Lorentz-frame. The boost-plane is spanned by the timelike vector $P^a= m(\partial_t)^a$ and its orthogonal spacelike counterpart $Q^a= m e^a$. Without loss of generality we may take $e^a$ to lie in the {\it x-z} plane \begin{align}\label{newvars} &m(e_z)^a = Q^a \cos\alpha - m\bar{e} ^a \sin\alpha \nonumber\\ & m(e_x)^a = Q^a \sin\alpha + m\bar{e} ^a \cos\alpha \end{align} where $\alpha$ denotes the angle between the axis of symmetry and the direction of the boost and $\bar{e} ^a$ the spacelike direction, which spans together with $(e_y)^a$ the two-plane orthogonal to the boost. With respect to (\ref{newvars}) the $\delta(z)$ factor in (\ref{KerrEM}) fixes $(Q x)= m\bar{x} \tan\alpha,\,\bar{x} = ( \bar{e} x) $, which in turn implies that \begin{equation}\label{radID} \rho^2 = x^2+y^2 = \frac{1}{m^2}( (Qx) \sin\alpha + m\bar{x} \cos\alpha )^2 + y^2= \frac{\bar{x} ^2}{\cos^2\alpha} + y^2 =: \bar{\rho}^2. \end{equation} So one ends up with the following expression for the energy-momentum tensor \begin{align}\label{covEM} T^a{}_b = & \frac{\delta(Qx-m\bar{x} \> \tan\alpha )}{8\pi\cos\alpha} \left\{ -\frac{2}{a}\left( \left[ \frac{a^2\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} ^3}\right ] -\frac{\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} } - \delta(\rhob- a) \right )P^a P_b \right. \nonumber\\ &+\left( 2\left [\frac{\bar{\rho}\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} ^3}\right ] -\frac{\pi}{a}\delta(\rhob- a) \right ) \left \{ P^a \>m (e_{\phi})_b + m (e_{\phi})^a\> P_b \right \}\nonumber\\ &+\frac{2}{a}\left( - \left[\frac{\bar{\rho}^2\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} ^3}\right ] - \frac{\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} } + 2\delta(\rhob- a) \right ) m (e_{\phi})^a\> m (e_{\phi})_b \nonumber\\ &\left. -\frac{2}{a}\frac{\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} } m (e_{\rho})^a\> m (e_{\rho})_b \right \} \end{align} where due to the $\delta$-factor in (\ref{covEM}) \begin{align} m (e_{\phi})^a &= \frac{1}{\rho}( x\, m (e_y)^a\! - y\, m (e_x)^a ) = \frac{1}{\bar{\rho}} \left( -y\sin\alpha\>Q^a\! - \frac{m}{\cos\alpha} (y\cos^2\alpha\>(\bar{e} )^a\! - \bar{x} \>(e_y)^a ) \right) \nonumber\\ m (e_{\rho})^a &= \frac{1}{\rho}(x\, m (e_x)^a\! + y \, m (e_y)^a) =\frac{1}{\bar{\rho}}\left( \bar{x} \tan\alpha\>Q^a\! + m(\bar{x} \>(\bar{e} )^a\! +y\>(e_y)^a ) \right) \end{align} Although this expression may look rather unwieldy in comparison with (\ref{KerrEM}) it allows a simple ultrarelativistic limit by letting $m\to 0$ and replacing $P^a$ and $Q^a$ by their null limit $p^a$. \begin{align}\label{urEM} T^a{}_b = & \frac{\delta(px)}{8\pi\cos\alpha} \left\{ -\frac{2}{a}\left( \left[ \frac{a^2\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} ^3}\right ] -\frac{\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} } - \delta(\rhob- a) \right ) \right. \nonumber\\ &+\left( 2\left [\frac{\bar{\rho}\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} ^3}\right ] -\frac{\pi}{a}\delta(\rhob- a) \right )\frac{2y\sin\alpha}{\bar{\rho}} \nonumber\\ &+\left( - \left[\frac{\bar{\rho}^2\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} ^3}\right ] + 2\delta(\rhob- a) \right )\frac{2y^2\sin^2\alpha}{a\bar{\rho}^2} \nonumber\\ &\left. -\frac{\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} }\frac{2\sin^2\alpha}{a} \right \}p^a p_b =: -\frac{1}{16\pi}g(\bar{x} ,y) \delta(px) \, p^a p_b \end{align} As expected the resulting energy-momentum tensor is that of a pp-(shock)wave. For $\alpha\to 0$ only the first term in the curly bracket survives and (\ref{urEM}) coincides with the result obtained in \cite{BaNa3}. \section*{\Large\bit 3) Perturbative evaluation} In order to find the metric corresponding to the distributional energy-momentum tensor (\ref{urEM}), one has to solve the Einstein equations that in this setting take the form of the Poisson equation \begin{equation}\label{Poisson} ( \partial_{\bar x}^2 + \partial_{y}^2 ) f(\bar x,y) = g(\bar x,y) \end{equation} for the profile-function $f(x)\, \delta(px)$ of the pp-wave. It can be solved straight-forwardly in a perturbative way except for the particular case of the orthogonal boost $\alpha\to\pi /2$ which needs special care and will therefore be dealt with in the next chapter. Rescaling $\bar{x} $ by $\cos\alpha$ and denoting the new variable by $x$ (\ref{Poisson}) becomes \begin{equation}\label{genPERT} (\Delta + \tan^2 \!\! \alpha \; \partial_x^2\; )\sum\limits_{n=0}^\infty f_n \sin^n\alpha =\frac{1}{\cos\alpha}(g_0 + \sin \alpha\: g_1 + \sin^2\! \alpha\: g_2) \end{equation} where the $g_i$ may be read off from (\ref{urEM}). Expanding $\cos\alpha$ and $\tan\alpha$ into power series with respect to $\sin\alpha$ and grouping corresponding powers together yields \begin{align} &\Delta f_0 = g_0,\qquad \Delta f_1 = g_1,\nonumber\\ &\Delta f_{2n} + \sum\limits_{k=0}^{n-1} \partial_x^2 f_{2k} = \frac{\poch{1/2}{n}}{n!} g_0+ \frac{\poch{1/2}{n-1}}{(n-1)!} g_2, \qquad n\geq 1,\nonumber\\ &\Delta f_{2n+1} + \sum\limits_{k=0}^{n-1} \partial_x^2 f_{2k+1}= \frac{\poch{1/2}{n}}{n!} g_1,\qquad n\geq 1. \end{align} Let us explicitly derive the first order perturbation $f_1$ which is determined by \begin{align}\label{pert} &\Delta f_1(x) = -8\frac{y}{\rho}\left( \left[\frac{\rho\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} ^3}\right ] -\frac{\pi}{2a}\delta(\rho- a) \right ). \end{align} In the region $0< \rho < a$ the classical analogue of (\ref{pert}) may be separated employing polar coordinates. Decomposing $f_1(x)$ into $\tilde{f}_1(\rho)\sin\phi$ we obtain the radial equation \begin{align} &\frac{1}{\rho}\partial_\rho(\rho\partial_\rho\tilde{f}_1 ) - \frac{1}{\rho^2}\tilde{f}_1 = -\frac{8\rho}{\sqrt{a^2-\rho^2} ^3}\nonumber\\ \intertext{which may be simplified by replacing $\rho$ by $a e ^u$}\nonumber &(\partial_u^2 -1)\tilde{f}_1 = -\frac{8e^{3u}}{\sqrt{1-e^{2u}}^3}. \end{align} This equation is easily solved by using the Green-function $$\vartheta(u)\sinh u = \frac{1}{2}\vartheta(\rho-a) \left( \frac{\rho}{a} -\frac{a}{\rho}\right ) $$ that gives rise to the particular solution $$ f_1(x)=\frac{8}{\rho}\sqrt{a^2-\rho^2} \sin\phi. $$ Taking into account the distributional identities \begin{align} \Delta\left( \frac{\vartheta(a -\rho) }{\rho}\sqrt{a^2-\rho^2} \sin\phi \right ) &= -\left[\frac{\rho\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} ^3}\right ]\sin\phi +\nonumber\\ &\hspace*{2cm}\frac{\pi}{2a}\delta(\rho- a) \sin\phi + 2\pi a\partial_y\delta^{(2)}(x)\nonumber\\ \Delta\left( \frac{1}{\rho}\sin\phi\right)&=2\pi \partial_y \delta^{(2)}(x), \end{align} we find \begin{equation}\label{perSOL} f_1(x) = \vartheta(a -\rho) \left(\frac{8}{\rho}(\sqrt{a^2-\rho^2} -a) \right ) \sin\phi - 8\vartheta(\rho-a) \, \frac{a}{\rho}\sin\phi. \end{equation} The choice of the solution is dictated by the distributional nature of the inhomogeneity and natural boundary conditions for $\rho\to\infty$ which ensure that $f_1$ vanishes asymptotically, which is equivalent to the fact that the whole solution tends to the AS-geometry at large distances. The last property may also be derived from the fact that $f_1$ tends to zero in the limit $a\to 0$. Comparing (\ref{perSOL}) with the result of the boost along the axis of symmetry \cite{BaNa3} shows that the rotational contributions being inverse powers die off only in the limit and not at some finite value. This behaviour is easily understood by interpreting (\ref{Poisson}) as a 2-dimensional electrostatic problem. In the spherically symmetric case the corresponding charge distribution produces only a monopole momentum and may therefore be replaced by a pointlike distribution of total charge outside its support. \section*{\Large\bit 4) Transversal Limit} In order to calculate the limit $\alpha\to \pi/2$ of (\ref{urEM}) one has to evaluate the expression on an arbitrary test function $\varphi$. Only two different types of expressions occur in the calculation, namely those which arise from the square-bracket terms and the concentrated (delta) contributions respectively. In the following we will exemplify the respective limits on typical representatives from each class. \begin{align}\label{example} &\lim_{\alpha\to \pi/2}\frac{1}{\cos\alpha} \left( \left[ \frac{\vartheta(a -\rhob) }{\sqrt{a^2-\rhob^2} ^3} \right ],\varphi \right ) = \lim_{\alpha\to \pi/2} \left( \left[ \frac{\vartheta(a -\rho) }{\sqrt{a^2-\rho^2} ^3} \right ],\tilde{\varphi} \right )= \nonumber\\ &\lim_{\alpha\to \pi/2}\int\limits_{\rho\leq a} d^2 x \frac{1}{\sqrt{a^2-\rho^2} ^3}\left( \varphi( \cos\alpha x,y) - \varphi( \cos\alpha a\cos\phi,a\sin\phi)\right)=\nonumber\\ &\int\limits_{\rho\leq a} d^2 x\frac{1}{\sqrt{a^2-\rho^2} ^3}\left( \varphi(0,y) - \varphi(0,a\sin\phi)\right), \end{align} where $\tilde{\varphi}(x,y):=\varphi(\cos\alpha x,y)$. Unfortunately the above result is not in a very useful form. Further simplification of (\ref{example}) may be achieved by integrating out $x$ in the first term, and $\rho$ in the second. However, in order to perform these integrations we have to restrict the domain of integration to a disk of radius $\bar{a} <a$ and do the limit $\bar{a} \to a$ in the end. More explicitely we find \begin{align*} &\int\limits_{\rho\leq\bar{a} } d^2 x\frac{1}{\sqrt{a^2-\rho^2} ^3} \varphi(0,y)= \int\limits_{-\bar{a} }^{\bar{a} } dy\varphi(0,y) \int\limits_{-\sqrt{\bar{a} ^2-y^2}}^{\sqrt{\bar{a} ^2-y^2} } dx \frac{1}{\sqrt{a^2-\rho^2} ^3}=\\ & \frac{2}{\sqrt{a^2-\bar{a} ^2}} \int\limits_{-\bar{a} }^{\bar{a} } dy \varphi(0,y) \frac{\sqrt{\bar{a} ^2-y^2 } }{a^2-y^2}\\ \intertext{and} &\int\limits_{\rho\leq\bar{a} } d^2 x\frac{1}{\sqrt{a^2-\rho^2} ^3} \varphi(0,a\sin\phi)= \int\limits_0^{\bar{a} } \frac{\rho d\rho}{\sqrt{a^2-\rho^2} ^3} \int\limits_0^{2\pi} d\phi \varphi(0,a\sin\phi)=\\ &\left(\frac{1}{\sqrt{a^2-\bar{a} ^2}}-\frac{1}{a} \right ) 2\int\limits_{-a}^a\frac{dy}{\sqrt{a^2-y^2} }\varphi(0,y). \end{align*} Using l'Hospital's rule (\ref{example}) becomes \begin{align*} &\frac{2}{a} \int\limits_{-a}^{a} \frac{dy}{\sqrt{a^2-y^2} }\varphi(0,y)- 2 \lim_{\bar{a} \to a} \sqrt{a^2-{\bar{a} }^2}\int\limits_{-\bar{a} }^{\bar{a} } \frac{dy}{\sqrt{\bar{a} ^2-y^2}(a^2-y^2)}\varphi(0,y)=\\ &\frac{2}{a}\int\limits_{-a}^{a} \frac{dy}{\sqrt{a^2-y^2} }\varphi(0,y) - \frac{\pi}{a}(\varphi (0,a) + \varphi (0,-a))=\\ &\left(\frac{2}{a}\frac{\vartheta(a-|y|)}{\sqrt{a^2-y^2} }\delta(\bar{x} ),\varphi\right ) - \frac{\pi}{a}\left(\delta(a-|y|) \delta(\bar{x} ) ,\varphi\right). \end{align*} The limit of the simplest concentrated contribution gives \begin{align}\label{example1} &\lim_{\alpha\to \pi /2}\frac{1}{\cos\alpha}\left( \delta(\rhob- a) ,\varphi\right)= \lim_{\alpha\to \pi /2} \left(\delta(\rho- a) ,\tilde{\varphi} \right)=\nonumber\\ &\lim_{\alpha\to \pi /2} a \int\limits_0^{2\pi} d\phi \varphi(\cos\alpha a\cos\phi,a\sin\phi)= a\int\limits_0^{2\pi} d\phi \varphi(0,a\sin\phi)=\nonumber\\ &2a\int\limits_{-a}^a \frac{dy}{\sqrt{a^2-y^2} }\varphi(0,y) = 2a\left(\frac{\vartheta(a-|y|)}{\sqrt{a^2-y^2} } \delta(\bar{x} ) ,\varphi \right) \end{align} Dealing with the remaining terms of (\ref{urEM}) in the same way we finally end up with the ``transversal'' energy-momentum tensor \begin{align}\label{orthEM} T^a{}_b= & \delta(px) \delta(\bar{x} )\left\{ \frac{1}{2}\delta(a-|y|) -\frac{1}{2}\left(\frac{y}{a}\right )\delta(a-|y|) \right\}p^ap_b = \nonumber\\ & \delta(px) \delta(\bar{x} )\delta(y+a)p^ap_b . \end{align} It is interesting to note that all contributions localized on the line segment $|y| \le a$ compensate each other and the energy-momentum tensor turns out to be concentrated on a pointlike region only. The corresponding profile function is obtained by solving the Poisson equation with (\ref{orthEM}) as inhomogeneity, which gives \begin{equation}\label{transPRO} f(x)= -8\log\left( \frac{\sqrt{\bar{x} ^2 + (y+a)^2} }{\rho_0}\right), \end{equation} where $\rho_0$ denotes the length-scale of the AS-geometry \cite{BaNa3}. The limit $a\to 0$ in fact reproduces the AS-profile function. \newpage \section*{\Large\bit 5) Conclusion} In the present paper we showed how to calculate the ultrarelativistic limit of the Kerr-geometry without putting any restriction on the direction of the boost. Our method is based upon the energy-momentum tensor of the original Kerr-geometry, which in contrast to metric admits a well-defined limit, thus avoiding possible ambiguities arising from the removal of the infinities of the metric-limit. The resulting energy-momentum tensor has the form of a pp-(shock)wave, depending parametrically on the angle $\alpha$ between the boost direction and the axis of rotation. This dependence turns the support of the energy-momentum tensor into an elliptical region in the two-dimensional subspace of the $px=0$-plane. Therefore only the limiting cases $\alpha=0$ and $\alpha=\pi/2$, where the support becomes a circle and a line-segment respectively, admit a solution in closed form. Nevertheless the general case allows a perturbative treatment if one suitably rescales the coordinates and expands the resulting expression with respect to $\sin\alpha$. An explicit calculation shows that the ''screening'' behaviour of the longitudinal ($\alpha=0$) case, where the solution turned into that of AS outside the disk with radius $a$, gets modified by contributions such that the whole solution displays only asymptotic AS-behaviour. The perturbative expansion breaks down in the limiting case $\alpha\to \pi/2$ where the direction of the boost becomes perpendicular to the axis of symmetry. Taking into account the distributional nature of the limit it is nevertheless possible to calculate the profile function of the limiting case in closed form. \par\noindent It would be interesting to investigate the dependence of particle scattering on the angle $\alpha$ in comparison to the $\alpha=0$ case, since the latter displays exactly the AS-behaviour outside the disk with radius $a$. Work in this direction is currently under progress. \newpage
\section*{Introduction} \addcontentsline{toc}{section}{\bf Introduction} Finding topological objects to characterize or even classify tilings could be seen as a major motivation to study tilings. This article is less ambitious but already has the taste of it in that it is largely devoted to the study of such an object: the integer group of coinvariants associated to the tiling and the range of a state on it. But it takes its stimulation from two areas of application. One is the $K$-theoretical gap labelling of Schr\"odinger operator s describing the motion of a particle in a tiling (and potentially many more features of solids for which a model by a tiling is appropriate), and the other concerns topological dynamical systems on the Cantor set. Let us briefly put these two areas into context. 1) Tilings furnish (discrete) models of solids which may not be periodic as for instance quasicrystals. A typical question arrising is that after the nature of the spectrum of a particle moving in the tiling (e.g.\ a phonon or an electron). The corresponding Schr\"odinger operator\ lies in the algebra of observables but its diagonalization is for non periodic tilings to difficult to carry out at present. To obtain at least some qualitative description of the spectrum Bellissard proposed the $K$-theoretical gap labelling \cite{Be4}. This gap labelling requires the computation of a topological invariant of a $C^*$-algebra\ which may be taken to be the algebra of observables, namely of the range of a tracial state on its $K_0$-group. The algebra of observables coincides with the algebra associated to the tiling and the tracial state comes from a trace on it. The $K_0$-group contains the integer group of coinvariants (at least for a large class of tilings) and the range of the tracial state equals its range on the coinvariants. This range is a countable subgroup of $\mbox{\rm \twlset R}$ which contains the possible labels of gaps in the spectrum of a Schr\"odinger operator, i.e.\ its elements are the gap labels predicted by $K$-theory. As it is determined by a topological invariant of the $C^*$-algebra\ associated to the tiling the $K$-theoretical gap labelling will be of use if the nature of the tiling is incorporated in the operator in a strong enough way so that all values of the above group (lying between $0$ and $1$) correspond to gaps. Although the quantification of the latter property is in general (in particular in higher dimensions) an unsolved problem we find it still worth persuing as there are no analytical alternatives in higher dimensions so far. 2) A certain class of tilings, the decorations of $\mbox{\rm \twlset Z}^d$, yield $d$-dimensional topological dynamical systems of the Cantor set. Such a topological dynamical system is given by $d$ commuting homeomorphisms, and a tiling which consists of decorated $d$-dimensional unit cubes touching nicely at their faces yields an example, the $d$ independent shifts yielding commuting homeomorphisms of the hull of the tiling. In one dimension the study of the $K$-group of the associated $C^*$-algebra\ (the one usually associated to the dynamical system coincides with the one associated to the tiling) turned out to be very fruitful. The ordered $K_0$-group with order unit, which coincides with the integer group of coinvariants as an unordered group, furnishes an invariant classifying the classes of strong orbit equivalent dynamical systems \cite{HPS,GPS}. In higher dimensions the interpretation of the $K$-group in terms of the properties of dynamical systems is yet less clear but it was recently realized that the (unordered) $K$-groups may be decomposed into cohomology groups of the group $\mbox{\rm \twlset Z}^d$ \cite{FoHu}. In particular the group of coinvariants furnishes part of the $K_0$-group thus motivating its computation which is carried out below for tilings which allow for an invertible substitution. In turn, most of the analysis done for decorations of $\mbox{\rm \twlset Z}^d$ remains valid for a much larger class of tilings including the Penrose tilings, as will be shown below using the concept of reduction.\bigskip Although the above applications are formulated in the framework of $K$-theory of $C^*$-algebra s we shall not emphasize the $C^*$-algebra ic aspects in this article since the topological invariant discussed here may be defined purely on the level of spaces, i.e.\ here of groupoids. The article is organized as follows. In the first section the groupoid associated to a tiling is described. We proceed along lines sligthly different from \cite{Ke2} by first introducing a multiplicative structure on the pattern classes of the tiling. This multiplicative structure is almost a groupoid, and using the notion of the radius of a pattern class one obtains the groupoid associated to the tiling in a way which makes its local nature transparent. In contrast to \cite{Ke2} this groupoid is only principal if the tiling is not periodic whereby we gain that it yields the right groupoid $C^*$-algebra\ in the periodic case, too. We then define the integer group of coinvariants associated to the tiling. Naturally, local manipulations, i.e.\ deformations of patterns, appear as maps which are close to being homomorphims of the introduced almost-groupoid s. In fact, a substitution\ can be algebraically defined as a particular kind of such a homomorphism. We discuss the concept of reduction and the notion of a decoration of $\mbox{\rm \twlset Z}^d$ and show that many tilings occuring in the description of quasicrystals reduce to decorations of $\mbox{\rm \twlset Z}^d$. Section 2 is devoted to the $C^*$-algebra ic characterization and contains some details about the $K$-theoretical gap labelling. We first recall the definition of the algebra associated to the tiling. The main result needed concerns tilings which reduce to decorations of $\mbox{\rm \twlset Z}^d$ and states that their $C^*$-algebra\ is stably isomorphic to a crossed product with $\mbox{\rm \twlset Z}^d$. In particular the $K$-theory of such a tiling is that of a topological dynamical system. The third section, which is kept independent of sections 1.4 and 2, is restricted to tilings which allow for an invertible substitution. Geometrically such a substitution\ may be obtained from an deflation but the analysis remains purely algebaic. The existence of an invertible substitution\ reflects in the possibility of obtaining an injective coding of the tiling in terms of the paths over a graph. This is technically the basis for the computation of the integer group of coinvariants and the range of its state. For one dimensional substitution s similar approaches have been successfully carried out, both, for the range of the state \cite{Q,BBG} as well as for the group itself \cite{For,Hos}. In higher dimensions results regarding the range of the state have been obtained under the restriction that the substitution\ forces its border \cite{Ke2}. The results obtained here are independent of the dimension of the tiling, the only requirements being that the substitution\ is (locally) invertible and, for the determination of the range of the state, primitive. In fact, we obtain a family of codings and graphs for a tiling which for substitution s which force their border all coincide and have the substitution\ matrix as connectivity matrix but are more complicated in the general case. As applications we discuss the Thue-Morse substitution\ sequence and the Penrose tilings. \section{Tilings and groupoids} Groupoids and almost-groupoid s play a central role in the description of tilings. They carry a structure which generalizes a group structure in so far as the multiplication of two elements $x$ and $y$ is only defined provided $(x,y)$ lie in a specified set. We shall indicate this by $x \vdash y$. As a general reference to groupoids which includes groupoid-$C^*$-algebra s we use \cite{Ren}. \subsection{Almost-groupoids} \begin{df} An almost-groupoid\ is a set $\Gamma$ together with a subset $\Gamma^{\vdash}\subset\Gamma\times\Gamma$ of composable elements (we write $x\vdash y$ for $(x,y)\in\Gamma^{\vdash}$) and a product map $x\vdash y\mapsto xy$: $\Gamma^{\vdash}\rightarrow\Gamma$ and an inversion map $x\mapsto x^{-1}$: $\Gamma\rightarrow\Gamma$, such that the following relations hold: \begin{itemize} \item[I1] $(x^{-1})^{-1} = x$, \item[I2] $x\vdash y$ implies $y^{-1}\vdash x^{-1}$, in which case $(xy)^{-1}=y^{-1}x^{-1}$, \item[I3] $x\vdash x^{-1}$ and $x\vdash x^{-1}x$ and $xx^{-1}x=x$, \item[A] $(x\vdash y$ and $xy\vdash z)$ whenever $(x\vdash yz$ and $y\vdash z)$, in which case $(xy)z = x(yz)$. \end{itemize} \end{df} The elements of $\Gamma^0=\{x^{-1}x|x\in\Gamma\}$ are called units. We will make frequent use of the maps $L,R:\Gamma\rightarrow\Gamma^0$ (often denoted $r,d$) defined by \begin{equation} L(x) = xx^{-1} \quad\quad R(x)=x^{-1} x. \end{equation} An almost-groupoid\ is called principal if the map $(L,R):\Gamma\to\Gamma^0\times\Gamma^0$ is injective. It defines orbits in the space of units: $u$ and $v$ are called orbit equivalent, $u\sim_{o} v$, whenever there is an $x$ such that $L(x)=u$ and $R(x)=v$. Note that $x\vdash y$ is equivalent to $x\vdash yy^{-1}y$ and $yy^{-1}\vdash y$ so that it is by $A$ equivalent to $x\vdash yy^{-1}$. In particular, $R(x)=L(y)$ implies $x\vdash y$. \begin{df} A groupoid is an almost-groupoid\ for which cancelation holds, i.e. \begin{itemize} \item[C] $x\vdash y$ and $x\vdash z$ imply $y= z$. \end{itemize} \end{df} This definition of a groupoid is equivalent to the usual one as e.g.\ presented in \cite{Ren}. It is not difficult to show that the cancelation axiom C indeed implies that \begin{itemize} \item $x\vdash y$ implies $R(x)=L(y)$, \item $x\vdash y$ and $y\vdash z$ imply $x\vdash z$. \end{itemize} In particular the above implications do not have to hold for almost-groupoid s and associativity is only guaranteed provided all multiplications are defined. The lack of cancelation may be put into an order. \begin{df} The order of an almost-groupoid\ is defined by \begin{equation} x\preceq y\quad\mbox{iff}\quad x\vdash y^{-1}\quad\mbox{and}\quad xy^{-1}=yy^{-1}. \end{equation} \end{df} A topological almost-groupoid\ is an almost-groupoid\ which carries a topology such that the product and the inversion map are continuous, $\Gamma^\vdash$ carrying the relative topology. A (locally compact) groupoid is called $r$-discrete if $\Gamma^0$ is open. A homomorphism of almost-groupoid s is a map which maps composable elements resp.\ units onto composable elements resp.\ units and preserves multiplication and inversion. In particular it preserves the order. A homomorphism of topological almost-groupoid s is a continuous homomorphism of almost-groupoid s. \begin{df} An inverse semi-group is an almost-groupoid\ for which $\Gamma^{\vdash}=\Gamma\times\Gamma$. \end{df} Any almost-groupoid\ can be made into an inverse semi-group by introducing an extra element $0$ and extending the multiplication to any $(x,y)\in\Gamma\times\Gamma$ by \begin{equation} x y = \left\{ \begin{array}{ll} xy & \mbox{if $x\vdash y$} \\ 0 & \mbox{else} \end{array} \right. \end{equation} and $0 x = x 0 = 0$, $0 0 = 0$. $0$ is its own inverse and greater than any other element. Conversely we could see an almost-groupoid\ as an inverse semi-group with an element $0$ satisfying $0 x = x 0 = 0$ for all $x$. However, homomorphisms of almost-groupoid s may in general not be extended to homomorphisms of inverse semi-groups and vice versa. The difference which makes homomorphisms of inverse semi groups unsuitable for the purposes of this article is that for a homomorphism $\varphi$ of an almost-groupoid, even if $\varphi(x)\vdash \varphi(y)$, it has only to satisfy $\varphi(x)\varphi(y)=\varphi(xy)$ if $x\vdash y$.\bigskip {\em Examples:} 1) The principal groupoid which will be associated to a tiling is given by an equivalence relation, i.e.\ its elements are pairs of equivalent elements of some set $X$. Multiplication is only defined for pairs $(x,y),(x',y')$ if $x'=y$ and then given by $(x,y)(y,z)=(x,z)$, and inversion by $(x,y)=(y,x)^{-1}$. The topology of the groupoids in question need in general not to coincide with the relative topology from $X\times X$. 2) Sometimes the equivalence may be expressed as orbit equivalence under the (right) action of a group $S$ acting on $X$: $x\sim y$ whenever $\exists s\in S:y=x\cdot s$. This leads to the consideration of another kind of groupoid which is called transformation group \cite{Ren}. Its space is the Cartesian product of $X$ with $S$ (here always considered to carry the product topology) and the groupoid structure is defined by $(x,s)(x',t)=(x,st)$ provided $x'=x\cdot s$ and $(x,s)^{-1}=(x\cdot s,s^{-1})$. We write it as $X\times_\alpha S$, $\alpha$ indicating the action. It may be viewed as the groupoid defined by orbit equivalence only if $S$ acts freely on $X$. 3) Any groupoid $\Gamma$ defines an inverse semi-group ${\cal ISG}(\Gamma)$ consisting of so-called $\Gamma$-sets. A $\Gamma$-set is a subset of $\Gamma$ for which the restrictions of $L$ and $R$ to it are both injective. If $X,Y$ are such $\Gamma$-sets (which may be empty) then $XY:=\{xy|x\in X,y\in Y,x\vdash y\}$ and $X^{-1}:=\{x^{-1}|x\in X\}$. This multiplication of sets is not only for $\Gamma$-sets defined. We will make frequent use of it. \subsection{Tilings} Borrowing the terminology for $d=2$ from \cite{GrSh} a $d$ dimensional tiling is a (countable) family of closed subsets of $\mbox{\rm \twlset R}^d$, called the tile s, which cover $\mbox{\rm \twlset R}^d$, overlap at most at their boundaries and may carry an additional decoration, e.g.\ to break symmetries. We restrict the possible shape of tile s through the requirements that they are of finite size, i.e\ they fit inside an $r$-ball for finite $r$, and they are the closure of their interior. We shall consider the geometrical objects as equivalence classes under translations in $\mbox{\rm \twlset R}^d$ (but not under rotations or reflections). A tiling or its class $T$ defines an almost-groupoid\ as follows. Consider the set of pattern class es (with resp.\ to translation) of $T$ where a pattern of a tiling is given by a finite subset of its tile s. A pattern class\ $M$ together with two tiles chosen from it (pointed out) will be called a doubly pointed pattern class. It will be denoted by $M_{xy}$ where $x$ is the first and $y$ the second pointed tile. The set of all doubly pointed pattern class es of $T$ will be denoted by $\mtxx{T}$ or simply by ${\cal M}_{\rm I\!I}$ if the tiling is clear from the context. ${\cal M}_{\rm I\!I}$ carries an order structure: $M_{x_1x_2}\dot{\preceq} N_{y_1y_2}$ iff the pattern class\ of $N_{y_1y_2}$ contains the pattern class\ of $M_{x_1x_2}$ at such a position that tile s $x_1$ resp.\ $x_2$ become $y_1$ resp.\ $y_2$, c.f.\ figure. \epsffile[0 0 430 110]{dFig6.ps} Moreover ${\cal M}_{\rm I\!I}$ carries the structure of an almost-groupoid\ whose order structure coincides with the one just introduced: Two elements $M_{x_1x_2}$, $N_{y_1y_2}$ are composable iff there is an $L_{z_1z_2}$ and a third tile $z$ of $L$ such that $M_{x_1x_2}\dot{\preceq} L_{z_1z}$ and $N_{y_1y_2}\dot{\preceq} L_{zz_2}$. We then define the product \begin{equation} M_{x_1x_2} N_{y_1y_2} = \min\{L_{z_1z_2}| \exists z\in\kt{L} :M_{x_1x_2}\dot{\preceq} L_{z_1z}, N_{y_1y_2}\dot{\preceq} L_{zz_2}\} \end{equation} where the minimum is taken with respect to $\dot{\preceq}$ and $\kt{L}$ denotes the tiles of the pattern class $L$. E.g.: \epsffile[0 0 430 120]{dFig5.ps} The inversion is given by \begin{equation} (M_{x_1x_2})^{-1} = M_{x_2x_1}. \end{equation} It is straightforward to see that $\dot{\preceq}$ equals the order relation of the almost-groupoid. So we will leave the dot away. We consider ${\cal M}_{\rm I\!I}$ as topological almost-groupoid\ carrying the discrete topology. The space of units ${{\mTxx}^0}$ is given by the elements of the form $M_{xx}$ which we simply write $M_x$ and call pointed pattern class es. The maps $L$ and $R$ act as \begin{equation} L(M_{xy})=M_{xx}\quad\quad R(M_{xy})=M_{yy}. \end{equation} Let $\kt{T}$ be the set of tile s of $T$ where we identify tile s if they are mapped onto each other by a translation which is a multiple of a period of the tiling. (Since $T$ is an equivalence class we are forced to do this.) In particular, for tilings which are completely periodic, i.e.\ periodic in $d$ independent directions, $\kt{T}$ is a finite set. We call a tiling together with a chosen tile a pointed tiling and denote it by $T_x$. Let $M_r(T_x)\in{{\mTxx}^0}$ be the smallest pointed pattern class\ which covers all balls of radius $r$ having their center inside tile\ $x$ and keep this tile chosen. Define $r_c(T_x,T_{x'}):=\sup\{r|M_r(T_x)=M_r(T_{x'})\}$. Then $d(T_x,T_{x'}):=e^{-r_c(T_x,T_{x'})}$ yields a metric on $\{T_x|x\in\kt{T}\}$ \cite{Ke2}. \begin{df} The hull $\Omega$ of $T$ is the completion of $\{T_x|x\in\kt{T}\}$ with respect to the tiling metric $d(T_x,T_{x'})=e^{-r_c(T_x,T_{x'})}$. \end{df} Note that $\Omega$ is a finite set if and only if $T$ is completely periodic. The elements of $\Omega$ may be interpreted as pointed tiling s since any of their patterns is eventually determined by some $T_x$. In other words $\omega\in\Omega$ may be thought of as being approximated by an increasing chain of pointed pattern class es (which we now write shorter $u$). \begin{df} A sequence of pointed pattern class es $\{u_\nu\}_\nu$ is called approximating if \begin{enumerate} \item any finite number of elements is composable, i.e.\ $\forall k\geq 1:u_1\dots u_k \vdash u_{k+1}$, \item $\mbox{\rm rad}(u_\nu)\to\infty$ where $\mbox{\rm rad}(u)$ is the largest $r$ such that all balls of radius $r$ having their center inside the pointed tile are covered by $u$. \end{enumerate} \end{df} An approximating sequence approximates a unique pointed tiling\ which we write as its limit, namely $\omega=\lim u_\nu$ is the pointed tiling\ which contains all pointed patterns $u_\nu$ at its pointed tile. It is clear that any pointed tiling\ has an approximation (just choose a divergent sequence $\{r_\nu\}_\nu$ in $\mbox{\rm \twlset R}^+$ and take $M_{r_\nu}(\omega)$) so that we may carry over the notion $\preceq$ by writing $u\preceq \omega$ if $u\preceq u_\nu$ for some element of an approximation of $\omega$. The pattern class es of $\omega$ are pattern class es of $T$ as well, i.e.\ for all $\omega\in\Omega$, $\mtxx{\omega}\subset {\cal M}_{\rm I\!I}$. One says that $\omega$ is locally homomorphic to $T_x$ (or $T$). Consequently $\omega$ (or its class) is called locally isomorphic to $\omega'$ if $\mtxx{\omega}=\mtxx{\omega'}$. $T$ is called minimal if it is locally isomorphic to any $\omega\in\Omega$ which is equivalent to saying that the closure of the orbit of any $\omega$ equals $\Omega$. Since the range of $d(T_x,\cdot)$ is discrete away from $0$ the hull is completely disconnected (zero dimensional). This means that the topology is generated by sets which are both open and closed. In fact, to any unit $u\in{{\mTxx}^0}$ we may assign an open and closed subset \begin{equation} U_u:=\{\omega\in\Omega|u\preceq \omega\} \end{equation} of $\Omega$. The collection of all of these yields a basis for the topology. Let ${\mTxx}_{2,\neq}$ denote the set of all doubly pointed pattern class es which are composed of two neighbored tiles (i.e.\ which have boundaries with nonzero intersection) and such that the first pointed tile is not equal to the second. ${\mTxx}_{2,\neq}$ generates ${\cal M}_{\rm I\!I}\backslash {\mTxx}_{1}$ where ${\mTxx}_{1}$ are the pointed pattern class es consisting of one tile only. The following compactness condition will be required furtheron. \begin{itemize} \item ${\mTxx}_{2,\neq}$ is a finite set. \end{itemize} Under this condition $\Omega$ is compact. In fact, this condition is equivalent to the requirement that for any $r>0$ there are only finitely many pointed pattern class es which are covered by a ball of radius $r$, and the latter has been used to prove compactness of $\Omega$ in \cite{Ke2}. To keep contact with \cite{Ke2} where proofs of the above statements may be found, we visualize a (non periodic) pointed tiling $T_x$ as a particular representative of $T$ in $\mbox{\rm \twlset R}^d$. For this we choose a reference point $0$ in $\mbox{\rm \twlset R}^d$ and introduce a puncture for each pattern class\ of a tile (i.e.\ the puncture is a point of the tile) so that $T_x$ is the representative for which tile\ $x$ lies with its puncture on $0$.\bigskip The almost-groupoid\ ${\cal M}_{\rm I\!I}$ acts on $\Omega$ from the right (and analogously from the left). Let \begin{equation} \Omega^\rhd=\{(\omega,c)|L(c)\preceq \omega\}\subset \Omega\times{\cal M}_{\rm I\!I} \end{equation} with relative topology. Define $\gamma:\Omega^\rhd\rightarrow \Omega:\, (\omega,c)\mapsto\omega\cdot c$ through an approximating sequence: Let $\lim u_\nu = \omega$ and set \begin{equation} (\lim u_\nu)\cdot c := \lim R(u_\nu c). \end{equation} More pictorially with $\omega=T'_x$ and $M_y\preceq T'_x$ we have $T'_x\cdot M_{yy'}:=T'_{x'}$ where $x'$ is the tile\ of $T'$ which coincides with $y'$ once $M_y$ has been realized as the pattern\ at $x$ in $T'$. Note that \begin{equation} \label{24071} L(c)\preceq \omega \mbox{ and } L(c')\preceq \omega\cdot c \quad\iff \quad c\vdash c'\mbox{ and } L(cc')\preceq \omega \end{equation} and that in this case \begin{equation}\label{24072} \gamma(\gamma(\omega,c),c')=\gamma(\omega,c c'). \end{equation} Moreover $\gamma(\cdot,M_{xy}):U_{M_x}\rightarrow U_{M_y}$ is continuous so that we might call it a continuous right action of an almost-groupoid. Let us define an equivalence relation on $\Omega^\rhd$ using approximations for the elements of $\Omega$: \begin{equation} (\lim u_\nu,c)\sim (\lim u_\nu,c')\quad\mbox{whenever}\quad \exists n: u_n c = u_n c'. \end{equation} It is straightforward to see that this definition is independent of the approximation and that the relation is transitive. We denote the equivalence classes by $[\omega,c]$. These classes may be visualized as doubly pointed tiling s, namely $[T'_x,M_{yz}]$ can be represented by $T'_{xz}$, $z$ being the tile\ of $T'_x$ corresponding to the $z$ in $M_{yz}$ once the latter has been identified in $T'_x$ such that $y$ coincides with $x$. \begin{lem}\label{l12041} Let ${\cal R}$ be quotient of $\Omega^\rhd$ by the above equivalence relation with quotient topology and consider the groupoid structure defined by \begin{equation}\label{06041} [\omega,c][\omega', c']=[\omega, c c'] \quad\mbox{provided}\quad \omega'=\omega\cdot c \end{equation} and \begin{equation} [\omega,c]^{-1}=[\omega\cdot c,c^{-1}]. \end{equation} Then ${\cal R}$ is an $r$-disrete groupoid. Its space of units is homeomorphic to $\Omega$. \end{lem} {\em Proof:} First, note that (\ref{06041}) is well defined because of (\ref{24071},\ref{24072}). It is straightforward to check that ${\cal R}$ satisfies the axioms of a groupoid. The topology of ${\cal R}$ is generated by sets of the form \begin{equation} {\cal U}_c:=\left[U_{L(c)}\times\{c\}\right]=\{[\omega,c]|L(c)\preceq\omega\}. \end{equation} ${\cal U}_c^{-1}={\cal U}_{c^{-1}}$ shows continuity of the inversion and $m^{-1}({\cal U}_c)=\bigcup_{c_1\vdash c_2, c\preceq c_1c_2}{\cal U}_{c_1}\times{\cal U}_{c_2}$ that of multiplication $m:{\cal R}^\vdash\to{\cal R}$. The space of units is $\{[\omega,u]|u=L(u)\preceq\omega\}$ which is open and homeomorphic to $\Omega$ via $[\omega,u]\mapsto \omega$.\hfill$\Box$ \begin{df} We call the groupoid ${\cal R}$ of Lemma~\ref{l12041} the groupoid associated to $T$. \end{df} The orbit of ${\cal R}$ through $\omega$ is given by $[\omega]_o=\{\omega\cdot c|L(c)\preceq \omega\}$ or more pictorial $[T'_x]_o=\{T'_x|x\in\kt{T'}\}$. Hence $[T'_x]_o$ is the equivalence class under translation $T'$: $T_x \sim_o T'_y$ whenever $T=T'$. If $T$ is not periodic then the quotient topology on the space of orbits $[\Omega]_o$ is not Hausdorff. Such a space is an example of a non commutative space in the sense of Connes \cite{Cone}. But the reader should be warned that what is called the non commutative space of Penrose tilings in \cite{Cone} does not agree with the space of orbits of the hull of a Penrose tiling but is rather a quotient of it. We shall comment on that in section~\ref{03081}. Let $\tilde{\gamma}:{\cal R}\rightarrow\Omega\times\Omega$ be given by $\tilde{\gamma}[\omega,c]=(\omega,\omega\cdot c)$. \begin{df} The principal groupoid ${\cal S}$ associated to $T$ is the image of ${\cal R}$ under $\tilde{\gamma}$ with weak topology and groupoid structure defined by the orbit equivalence under ${\cal R}$. \end{df} \begin{lem} ${\cal S}$ is isomorphic to ${\cal R}$ if and only if $T$ is non periodic. \end{lem} {\em Proof:} Clearly $\tilde{\gamma}$ is a surjective and open homomorphism of topological groupoids. But it is only injective if $\gamma$ is free in the sense that $\omega\cdot c=\omega$ implies that $c$ is a unit. This is precisely the case if $T$ is not invariant under a translation.\hfill$\Box$\bigskip In particular ${\cal R}$ is principal if $T$ has no periodicity. Note that the topology of ${\cal S}$ does not coincide with the relative topology inherited from $\Omega\times\Omega$. Its topology is generated by \begin{equation} U_{M_{xy}}:=\tilde{\gamma}[{\cal U}_{M_{xy}}]=\{(T'_{x'},T'_{y'})|M_x\preceq T'_{x'}, T'_{y'}=T'_{x'}\cdot M_{xy}\}. \end{equation} In \cite{Ke2} the principal groupoid of a tiling was used instead of ${\cal R}$ but it turned out that from the point of view of physics the groupoid-$C^*$-algebra\ of ${\cal S}$ is not appropriate for tilings which have a periodicity in that it does not contain all translation operators in that case. \subsubsection{The integer group of coinvariants associated to a tiling} An important ingredient for the characterization of a tiling is the integer group of coinvariants of its associated groupoid. This notion is borrowed from the case of transformation groups $X\times_\alpha S$ where the action of $S$ on $X$ gives rise to an action on the group $C(X,\mbox{\rm \twlset Z})$ of continuous integer-valued functions, $\alpha(s)^*(f)(x)=f(x\cdot s )$. The coinvariants are then the quotient of $C(X,\mbox{\rm \twlset Z})$ by the subgroup generated by $\{f-\alpha(s)^*(f)|f\in C(X,\mbox{\rm \twlset Z}),s\in S\}$.\bigskip The ample semi group ${\cal ASG}(\Gamma)$ of an $r$-discrete groupoid $\Gamma$ is the sub-inverse-semi-group of ${\cal ISG}(\Gamma)$ given by its compact open $\Gamma$-sets. ${\cal ASG}(\Gamma)$ acts on ${\cal ASG}(\Gamma)^0$ by conjugation, $U\cdot {\cal U} := {\cal U}^{-1} U{\cal U}$, leading to a partially defined action on $\Gamma^0$ \cite{Ren}: for $u\in L({\cal U})$, \begin{equation} \{u\}\cdot {\cal U}={\cal U}^{-1}\{u\}\,{\cal U}=\{c^{-1}uc\} \end{equation} where $c=L^{-1}(u)$. The integer group of coinvariants of $\Gamma$ (or group of coinvariants with coefficients in $\mbox{\rm \twlset Z}$) shall be the coinvariants of the $\mbox{\rm \twlset Z}$-module $C_c(\Gamma^0,\mbox{\rm \twlset Z})$ of integer valued continuous functions over $\Gamma^0$ with compact support with respect to the above partially defined action of ${\cal ASG}(\Gamma)$. More precisely define $\eta:{\cal ASG}(\Gamma)\to C_c(\Gamma^0,\mbox{\rm \twlset Z})$ by \begin{equation} \label{23061} \eta({\cal U}) := \chi^{}_{L({\cal U})}-\chi^{}_{R({\cal U})} \end{equation} where $\chi^{}_{U}$ is the characteristic function on $U$ and let $E_\Gamma$ be the subgroup of $C_c(\Gamma^0,\mbox{\rm \twlset Z})$ generated by $\mbox{\rm im}\,\eta$. \begin{df} The integer group of coinvariants of an $r$-discrete groupoid $\Gamma$ is $$H(\Gamma)=C_c(\Gamma^0,\mbox{\rm \twlset Z})/E_\Gamma.$$ \end{df} If ${\cal U}$, ${\cal U}'$ are disjoint $\Gamma$-sets and ${\cal U}\cup{\cal U}'$ is a $\Gamma$-set as well then $\eta({\cal U}\cup {\cal U}') = \eta({\cal U}) +\eta({\cal U}')$. For transformation groups ${\cal ASG}(X\times_\alpha S)$ contains disjoint unions of sets of the form $U\times\{s\}$ where $s\in S$ and $U\subset X$ is open and compact. Since $\eta(U\times\{s\}) = \chi^{}_U-\chi^{}_{U\cdot s}$ the above definition is indeed a generalization. Let us consider groupoids associated to tilings. Since the restrictions of $L$ and of $R$ to the sets ${\cal U}_c$ generating the topology are both injective ${\cal ASG}({\cal R})$ contains unions of these sets. These unions may be chosen to be disjoined because the hull is zero dimensional. Now using $L({\cal U}_c)={\cal U}_{L(c)}$ one obtains \begin{equation} H({\cal R}) = \langle\{\chi^{}_{U_u}|u\in {{\mTxx}^0}\}\rangle/ \langle\{\chi^{}_{U_{L(c)}}-\chi^{}_{U_{R(c)}}|c\in {\cal M}_{\rm I\!I}\}\rangle. \end{equation} We call $H({\cal R})$ also the integer group of coinvariants associated to $T$. \subsection{Reduction} The concept of reduction is very important in the sequel. In the context of dynamical systems it is sometimes called induction but we prefer the notion established for groupoids \cite{Ram}.\bigskip Let ${\cal N}$ be a sub-almost-groupoid\ of ${\cal M}_{\rm I\!I}$ and define \begin{equation} \Omega_{\cal N} := \bigcup_{c\in{\cal N}} U_{R(c)} \end{equation} as well as \begin{equation} \Omega^\rhd_{\cal N} := \Omega^\rhd\cap(\Omega_{\cal N}\times {\cal N}) \end{equation} with relative topologies. We shall often consider the case in which ${\cal N}$ is finitely generated, i.e.\ ${\cal N}= \langle{\cal C}\rangle$ for a finite subset ${\cal C}$ of ${\cal M}_{\rm I\!I}$, then $\Omega_{\cal N} = \bigcup_{c\in{\cal C}} (U_{R(c)}\cup U_{L(c)})$. In fact, since $\Omega$ is compact and the $U_{u}$ form a basis of its topology any open and closed subset of $\Omega$ arrises in this manner with a finitely generated ${\cal N}$. There are two groupoids which naturally arrise from such a situation, one is the so-called reduction of ${\cal R}$ by $\Omega_{\cal N}$: \begin{equation} {\cal R}_{\Omega_{\cal N}}:=\{[\omega,c]\in{\cal R}|\,\omega,\omega\cdot c\in\Omega_{\cal N}\} \end{equation} (with relative topology), and the other is \begin{equation} {\cal R}_{\cal N}:=\left[\Omega_{\cal N}^\rhd\right]_{\cal N} \end{equation} where $[\omega,c]_{\cal N}=\{(\omega,c')\in [\omega,c]|c'\in{\cal N}\}$ (with quotient topology). \begin{df} We call ${\cal N}$ approximating if any $\omega\in\Omega_{\cal N}$ is approximated by a sequence $\{u_\nu\}_\nu$ with elements in ${\cal N}$ and generating if $[\omega,c]_{\cal N}\mapsto [\omega,c]:\,{\cal R}_{\cal N}\to{\cal R}_{\Omega_{\cal N}}$ is an isomorphism. We call ${\cal N}$ regular if any orbit of ${\cal R}$ intersects $\Omega_{\cal N}$, i.e.\ if $[{\cal R}]_o=[{\cal R}_{\cal N}]_o$. \end{df} Note that $[\omega,c]_{\cal N}\mapsto [\omega,c]:\,{\cal R}_{\cal N}\to{\cal R}_{\Omega_{\cal N}}$ provides an isomorphism if it is surjective. Note also that for minimal $T$ any non-empty ${\cal C}\in{\cal M}_{\rm I\!I}$ generates a regular sub-almost-groupoid\ $\erz{\C}$. None of the properties of the definition implies any other as counterexamples show. However, if ${\cal N}$ is generating then $\tilde{{\cal N}}=\{c_1uc_2|c_i\in{\cal N},c_1\vdash c_2,R(c_1),L(c_2)\preceq u\in{{\mTxx}^0}\}$ is approximating and ${\cal R}_{\cal N}={\cal R}_{\tilde{{\cal N}}}$. \begin{lem} If ${\cal N}$ is approximating then the topology of $\Omega_{\cal N}$ is generated by sets of the form $U_u$ with $u\in{\cal N}^0$. \end{lem} {\em Proof:} It is clear that the topology of $\Omega_{\cal N}$ is generated by sets of the form $U_{uv}$ where $u\in {\cal N}^0$ and $u\vdash v\in{{\mTxx}^0}$. If ${\cal N}$ is approximating then there exists for any $\omega\in U_{uv}$ a $u_\omega\in{\cal N}^0$ such that $uv\preceq u_\omega\preceq\omega$. Hence $U_{uv}=\bigcup_{uv\preceq\omega} U_{u_\omega}$. \hfill$\Box$ \begin{lem}\label{18071} If ${\cal N}$ is regular then for any $u\in{{\mTxx}^0}$ there is a finite subset $\{c_i\}_i\subset{\cal M}_{\rm I\!I}$ with $U_u=\dot{\bigcup}_i U_{L(c_i)}$ and $U_{R(c_i)}\subset\Omega_{\cal N}$. \end{lem} {\em Proof:} By regularity we may find for $\omega\in U_u$ a $c\in{\cal M}_{\rm I\!I}$ such that $\omega\cdot c\in\Omega_{\cal N}$. Let $r=\sup_{\omega\in U_u} \inf\{r'|u\preceq M_{r'}(\omega),\exists c\in{\cal M}_{\rm I\!I}:\omega\cdot c\in\Omega_{\cal N},L(c)\preceq M_{r'}(\omega)\}$. Since $U_u$ is compact $r$ is finite. Choose for any $\omega\in U_u$ a $c_\omega$ so that $\omega\cdot c_\omega\in\Omega_{\cal N}$ and $L(c_\omega)=M_r(\omega)$. In particular $U_{R(c_\omega)}\subset\Omega_{\cal N}$ and $\omega\in U_{L(c_\omega)}\subset U_u$. Hence $U_u=\bigcup_{u\preceq\omega}U_{L(c_\omega)}$. Since $U_{M_r(\omega)}\cap U_{M_r(\omega')}$ is either $U_{M_r(\omega)}$ or empty the union can be made disjoint and then, by compactness, finite. \hfill$\Box$\bigskip The embedding $i:{\cal R}_{\Omega_{\cal N}}\hookrightarrow {\cal R}$ furnishes a surjective group homomorphism $i^*:C(\Omega,\mbox{\rm \twlset Z})\to C(\Omega_{\cal N},\mbox{\rm \twlset Z})$ which satisfies $i^*\circ\eta=\eta\circ i^{-1}$. Therefore $E_{{\cal R}_{\Omega_{\cal N}}}=E_{\cal R}\cap C(\Omega_{\cal N},\mbox{\rm \twlset Z})$ and $i^*$ induces a surjective homomorphism $H({\cal R})\to H({\cal R}_{\Omega_{\cal N}})$. \begin{lem} \label{23051} If ${\cal N}$ is regular then $H({\cal R})=H({\cal R}_{\Omega_{\cal N}})$. \end{lem} {\em Proof:} We have to show that that for $u\in{{\mTxx}^0}$ there is a finite $\{u_i\}_i\subset{{\mTxx}^0}$ with $U_{u_i} \subset\Omega_{\cal N}$ and $[\chi^{}_u]_{E_{\cal R}}=\sum_i [\chi^{}_{u_i}]_{E_{{\cal R}}}$. This follows immediately from the foregoing lemma taking $u_i=R(c_i)$. \hfill$\Box$\bigskip {\em Remark:} Lemma~\ref{18071} implies even that the groupoids ${\cal R}$ and ${\cal R}_{\Omega_{\cal N}}$ are continuously similar: Let ${\cal C}$ be the collection of $c_i$'s such that $\Omega=\dot{\bigcup}_i U_{L(c_i)}$ and $U_{R(c_i)}\subset\Omega_{\cal N}$. Define $\Omega\to{\cal C}:\omega\mapsto c_\omega$, $c_\omega$ being the unique element satisfying $L(c_\omega)\preceq \omega$, and $\epsilon:{\cal R}\to{\cal R}_{\Omega_{\cal N}}$: \begin{equation} \epsilon[\omega,c]=[\omega\cdot c_\omega,c_\omega^{-1}cc_{\omega\cdot c}]. \end{equation} Clearly $\epsilon$ is continuous and $\epsilon\circ i = id_{{\cal R}_{\Omega_{\cal N}}}$ for the embedding $i:{\cal R}_{\Omega_{\cal N}}\to {\cal R}$. But also $i\circ\epsilon$ is similar to the identity on ${\cal R}$, namely the map $\Theta:\Omega\to{\cal R}$: $\Theta(\omega)=[\omega,c_\omega]$ satifies $\Theta(L([\omega,c]))\epsilon([\omega,c])=[\omega,c]\Theta(R([\omega,c]))$. This is a continuous version of the theorem of Ramsay \cite{Ram} guaranteeing that the continuous cohomology of ${\cal R}_{\Omega_{\cal N}}$ is isomorphic to that of ${\cal R}$ \cite{Ren}. \subsection{Locally defined maps between tilings}\label{11071} We want to formulate in the present framework the concept of locally deriving one tiling from another, a concept which has been introduced for tilings in \cite{BSJ}. Roughly speaking a local derivation amounts to a deformation of pattern class es, e.g.\ by adding or deleting or deforming boundaries or decorations of tiles. Such a deformation will be given by a certain class of maps from sub-almost-groupoid s of ${\cal M}_{\rm I\!I}$ to the doubly pointed pattern class es ${\cal M}_{\rm I\!I}'=\mtxx{T'}$ of the other tiling class $T'$. The results will be partly needed in the third section. \begin{lem}\label{27061} Let ${\cal N}$ be an approximating sub-almost-groupoid\ of ${\cal M}_{\rm I\!I}$ and $\hat{\varphi}:{\cal N}\to{\cal M}_{\rm I\!I}'$ be a map which preserves composability, commutes with the inverse map and satisfies \begin{itemize} \item[1] $\hat{\varphi}(c)\hat{\varphi}(c') \preceq \hat{\varphi}(cc')$ \item[2] (growth condition) there is a $t>0$ such that $|\mbox{\rm rad}(\hat{\varphi}(u))-t\,\mbox{\rm rad}(u)|$ is a bounded function on ${{\mTxx}^0}$. \end{itemize} Then \begin{equation} \varphi[\lim u_\nu,c]_{\cal N} := [\lim\hat{\varphi}(u_\nu),\hat{\varphi}(c)] \end{equation} defines a homomorphism $\varphi:{\cal R}_{\cal N}\rightarrow{\cal R}'$. \end{lem} {\em Proof:} Let us first show that $\svarphi$ preserves the order. $c\preceq c'$ is by definition $c\vdash {c'}^{-1}$ and $c'=cR(c')$. But then $\svarphi(c)\preceq \svarphi(c)\svarphi(R(c'))\preceq \svarphi(c')$. Now let $\{u_\nu\}_\nu$ be an approximating sequence. Then for any $1<k$, $u_1\dots u_k\vdash u_{k+1}$. The first condition and preservation of composability immediatly implies that $\svarphi(u_1)\dots \svarphi(u_k)\vdash \svarphi(u_{k+1})$. By the growth condition the radius of $\{\svarphi(u_\nu)\}_\nu$ diverges so it is an approximating sequence as well. Hence for $\omega=\lim u_\nu$, $\varphi(\omega)=\lim \hat{\varphi}(u_\nu)$ is well defined and clearly continuous. As well is $\Omega_{\cal N}\times {\cal N}\to \Omega'\times{\cal M}_{\rm I\!I}'$: $(\omega,c)\mapsto (\varphi(\omega),\hat{\varphi}(c))$. Thus we are left to show that $\varphi$ induces a map on $[\Omega_{\cal N}^\rhd]_{\cal N}$ and preserves the groupoid structure. (By definition of the topology its continuity is then guaranteed.) We have $\svarphi(L(c))\succeq L(\svarphi(c))$ and hence $\svarphi(R(u_\nu c))\succeq R(\svarphi(u_\nu)\svarphi(c))$. Therefore does $L(c)\preceq\omega$ imply $\svarphi(L(c))\preceq\varphi(\omega)$ and $\varphi(\omega\cdot c)=\lim\svarphi(R(u_\nu c))=\lim R(\svarphi(u_\nu)\svarphi(c))= \varphi(\omega)\cdot\svarphi(c)$. This shows that $\varphi(\Omega_{\cal N}^\rhd)\subset{\Omega'}^\rhd$ and that $\varphi$ maps equivalent elements onto equivalent ones, because equivalence may be expressed as $(\omega,c)\sim (\omega,c')$ whenever $\exists c'':c,c'\preceq c''$ and $L(c'')\preceq\omega$. That $\varphi$ commutes with inversion on the level of groupoids follows from the corresponding hypothesis on the level of the almost-groupoid s. Finally, multiplication is preserved if $(\varphi(\omega),\svarphi(c)\svarphi(c'))\sim (\varphi(\omega),\svarphi(cc'))$. In view of the above characterization of the equivalence this follows from $\hat{\varphi}(c)\hat{\varphi}(c')\preceq\hat{\varphi}(cc')$. \hfill$\Box$\bigskip We say that $\varphi:{\cal R}_{\cal N}\rightarrow{\cal R}'$ is locally defined. The corresponding homomorphism on the principal groupoids is simply $(\varphi,\varphi):{\cal S}_{\cal N}:={\cal S}\cap(\Omega_{\cal N}\times\Omega_{\cal N})\to{\cal S}'$. Details of the proof of the following theorem, which is constructive, will be used in section 3. \begin{thm}\label{25071} Let $\varphi:{\cal R}\to{\cal R}'$ be a locally defined homomorphism. If $\varphi$ has a left inverse then this left inverse is locally defined. \end{thm} {\em Proof:} Let $a\in{\mTxx}_{1}$, ${\mTxx}_{1}$ denoting the set of pointed pattern class es of tiles. Since $\varphi(U_a)$ is open and compact we may find a finite subset $\Phi(a)$ of ${\cal M}_{\rm I\!I}'$ satisfying \begin{equation} \label{26061} \varphi(U_a)=\bigcup_{v\in \Phi(a)}U_{v}. \end{equation} Let $E^0(u)=\{e\in{\cal M}_{\rm I\!I}|R(e)=u\}$ and denote by $x(e)$ the first pointed tile of $e$. The expression $u=\prod_{e\in E^0(u)} e^{-1}x(e)e$ motivates to define \begin{equation} \Phi^e(u):=\{\hat{\varphi}(e^{-1})\}\Phi(x(e))\{\hat{\varphi}(e)\} \end{equation} where we have used set multiplication, and \begin{equation} \Phi(c):=\left(\prod_{e\in E^0(L(c))}\Phi^e(L(c))\right)\{\hat{\varphi}(c)\}. \end{equation} We claim that \begin{equation}\label{27062} \varphi({\cal U}_c) = \bigcup_{d\in\Phi(c)}{\cal U}_d. \end{equation} Indeed, using that $u\preceq\omega$ whenever $\forall e\in E^0(u): x(e)\preceq\omega\cdot e^{-1}$ we get \begin{eqnarray*} \varphi({\cal U}_c) & = & \bigcap_{e\in E^0(L(c))} \{[\varphi(\omega),\hat{\varphi}(c)]|\omega\cdot e^{-1}\in U_{x(e)}\} \\ & = & \{[\omega',\hat{\varphi}(c)]| \forall e\in E^0(L(c)): \omega'\in \bigcup_{v\in\Phi(x(e))}U_{\hat{\varphi}(e^{-1})v\hat{\varphi}(e)}\} \\ &= & \{[\omega',\hat{\varphi}(c)]|\omega'\in \bigcup_{v\in\Phi(L(c))}U_v\} \end{eqnarray*} from which the claim follows. In particular, setting ${\cal N}=\Phi({\cal M}_{\rm I\!I})$, we have $\mbox{\rm im}\,\varphi={\cal R}'_{\cal N}$. Furthermore let $\Phi(c)\vdash \Phi(c')$, which shall mean that $\exists d\in\Phi(c)\,\exists d'\in\Phi(c'):d\vdash d'$. Then from (\ref{27062}) we may conclude that $U_{R(c)}\cap U_{L(c)}\neq\emptyset$. Hence \begin{equation}\label{27063} \Phi(c)\vdash \Phi(c')\quad\mbox{implies}\quad c\vdash c'. \end{equation} Now set $W(d)=\{c\in{\cal M}_{\rm I\!I}|d\in\Phi(c)\}$ which by construction is finite. If $c,c'\in W(d)$ then by (\ref{27062}) ${\cal U}_c\cap{\cal U}_{c'}\neq\emptyset$ and hence $c$ and $c'$ have a common greater element, e.g.\ $L(c)c'$. Since $L(d)\in\Phi(L(c))\cap \Phi(L(c'))$ also $L(d)\in\Phi(L(c)L(c'))$ so that $d\in\Phi(L(c)c')$. It follows that $\min\{c|\forall c'\in W(d):c'\preceq c\}=\max W(d)$. Define $\hat{\psi}:{\cal N}\to{\cal M}_{\rm I\!I}$: \begin{equation} \hat{\psi}(d):=\max W(d). \end{equation} It commutes with the inverse. Because of $d\in\Phi(\hat{\psi}(d))$ and (\ref{27063}) $d\vdash d'$ implies $\hat{\psi}(d)\vdash \hat{\psi}(d')$. But it also implies $W(d)W(d')\subset W(dd')$ and hence $\hat{\psi}(d)\hat{\psi}(d')\preceq\hat{\psi}(dd')$. Since $\Phi(A)$ is finite $\hat{\psi}$ and $\hat{\psi}\circ\hat{\varphi}$ satisfy the growth condition (with $t^{-1}$ resp.\ $1$ in place of $t$) and ${\cal N}$ is approximating. Hence they extend to a homomorphisms ${\cal R}'_{\cal N}\to {\cal R}$ resp.\ ${\cal R}\to {\cal R}$. Since $\psi\circ\varphi$ extends to the identity $\psi$ extends to a left inverse.\hfill$\Box$\bigskip We say that $\varphi$ is locally invertible (on its image). \begin{lem} If $T$ is non-periodic then the almost-groupoid\ ${\cal N}=\Phi({\cal M}_{\rm I\!I})$ occurring in the proof of the above theorem is generating. \end{lem} {\em Proof:} Consider the commuting diagram \begin{eqnarray*} {\cal R} & \stackrel{\varphi}{\leftrightarrow} & {\cal R}'_{{\cal N}}\\ \tilde{\gamma}\updownarrow & & \downarrow \tilde{\gamma}' \\ {\cal S}& \stackrel{(\varphi,\varphi)}{\to} & {\cal S}'_{{\cal N}} \end{eqnarray*} where $\tilde{\gamma}$ is an isomorphism since $T$ is non periodic. Clearly $(\varphi,\varphi)$ is injective. Suppose that it is not surjective. Then there must be an $x\sim_o y$ such that $\varphi^{-1}(x)\not\sim_o \varphi^{-1}(y)$. But this contradicts that $\varphi^{-1}$ is a left inverse which preserves (like any homomorphism of groupoids) orbits. Therefore $\tilde{\gamma}'$ is bijective and hence ${\cal N}$ generating.\hfill$\Box$\bigskip Counterexamples show that the requirement for $T$ to be non periodic is neccessary. \subsection{Decorations of $\mbox{\rm \twlset Z}^d$} For many tilings the groupoid is a transformation group the group being $\mbox{\rm \twlset Z}^d$. This is e.g.\ the case if the tiling is built from unit cubes which nicely touch at the faces. But since we allow for decorat ions these tilings do not have to be periodic. They are called decorations of $\mbox{\rm \twlset Z}^d$. In any case they allow for $d$ commuting homeomorphisms of the hull which are given by the $d$ independent shifts such that the orbits of ${\cal R}$ become orbits under the action of $\mbox{\rm \twlset Z}^d$ and ${\cal R}$ a transformation group $\Omega\times_\alpha\mbox{\rm \twlset Z}^d$. The structure of the $K$-groups of $C^*$-algebra s defined by such transformation groups is fairly well understood and it turns out to be useful as well in the case where ${\cal R}$ merely has a reduction ${\cal R}_{\Omega_{\cal N}}$ which is isomorphic to $\Omega_{\cal N}\times_\alpha\mbox{\rm \twlset Z}^d$. In view of Lemma~\ref{23051} this already implies for regular ${\cal N}$ that the integer group of coinvariants coincide. \begin{df} We say that $T$ reduces to a decoration of $\mbox{\rm \twlset Z}^d$ if there is a regular ${\cal N}$ such that ${\cal R}_{\Omega_{\cal N}}\cong\Omega_{\cal N}\times_\alpha\mbox{\rm \twlset Z}^d$. If ${\cal N}={\cal M}_{\rm I\!I}$ then $T$ is called a decoration of $\mbox{\rm \twlset Z}^d$. \end{df} Tilings which reduce to decorations of $\mbox{\rm \twlset Z}^d$ yield by definition $d$-dimensional topological dynamical systems. The topological space is $\Omega_{\cal N}$ which is homeomorphic to the Cantor set. These systems are topologically transitive and the notion of minimality in the sense of tilings and in the sense of dynamical systems coincide.\medskip There is a large class of tilings which reduce to decorations of $\mbox{\rm \twlset Z}^d$. These tilings are composed of $d$ dimensional parallel epipeds which may be decorated. For their definition consider a set of $N$ vectors $\xi_1,\cdots,\xi_N$ which span $\mbox{\rm \twlset R}^d$. A set $J\subset \{1,\cdots,N\}$ containing $n$ elements, $n\leq d$, such that $\{\xi_i\}_{i\in J}$ are linear independent defines a subset $\{\sum_{i\in J}c_i\xi_i|c_i\in [0,1]\}$ of $\mbox{\rm \twlset R}^d$ which is an $n$ dimensional parallel epiped. We will call a translate of it an $n$-facet of type $J$. Their boundaries are $n'$-facets, $n'<n$, of type $J'\subset J$. Let $T$ be a tiling of $\mbox{\rm \twlset R}^d$ consisting of possibly decorat ed \pe{d}s which are arranged in such a way that \begin{itemize} \item[D1] the boundaries of tile s overlap, if at all, at common complete $d'$-facets, $d'<d$. \end{itemize} Provided the number of different decorat ions is finite, which we assume, there is only a finite number of pattern class es consisting of two tiles which touch each other so that the hull is compact. If $d=N$ then $T$ itself a decoration of $\mbox{\rm \twlset Z}^d$ so here we are interested in $d<N$. Now fix a type $I_0$ of some tile\ appearing in $T$ and set \begin{equation}\label{19061} {\cal C}=\{a\in{\mTxx}_{1}|a\mbox{ is of type }I_0\}. \end{equation} where ${\mTxx}_{1}$ denotes the set of pointed pattern class es of tiles. For any $i\in I_0$ choose a (common) normal for $d\!-\!1$-facets of type ${I_0\backslash\{i\}}$, i.e.\ a positive and a negative side, and let ${\cal B}_i\subset{\mTxx}_{2,\neq}$ consist of those doubly pointed pattern class es which have as common boundary a \pe{d-1}\ of type ${I_0\backslash\{i\}}$ and such that their first resp.\ second pointed tile is on the positive resp.\ negative side of the common boundary. Then $\Omega_{\erz{\C}}=\bigcap_i\Omega_{\erz{{\cal B}_i}}$. Define $\breve{\alpha}_i:\Omega_{\erz{{\cal B}_i}}\to\Omega_{\erz{{\cal B}_i}}$ by \begin{equation} \breve{\alpha}_i(\omega) = \omega\cdot c \end{equation} where $c$ is the unique element of ${\cal B}_i$ for which $L(c)\preceq\omega$. Clearly $\breve{\alpha}^{-1}_i(\omega) = \omega\cdot c^{-1}$ with the unique $c$ for which $R(c)\preceq\omega$. Now we require in addition that \begin{itemize} \item[D2] for all $i\in I_0$ and all $\omega\in \Omega_{\erz{\C}}$ there are natural numbers $n,m>0$ such that $\breve{\alpha}_i^n(\omega)\in \Omega_{\erz{\C}}$ and $\breve{\alpha}_i^{-m}(\omega)\in \Omega_{\erz{\C}}$. \end{itemize} Let $n^\pm_i(\omega)$ be the smallest $n>0$ for which $\breve{\alpha}_i^{\pm n}(\omega)\in \Omega_{\erz{\C}}$. This defines maps $n^\pm_i:\Omega_{\erz{\C}}\to\mbox{\rm \twlset Z}^{>0}$ which are continuous by the compactness of $\Omega_{\erz{\C}}$. Consequently $\alpha:\Omega_{\erz{\C}}\to\Omega_{\erz{\C}}$ \begin{equation} \alpha_i(\omega):=\breve{\alpha}_i^{n^+_{i}(\omega)}(\omega) \end{equation} are continuous maps (first return maps) having inverses $\alpha_i^{-1}:=\breve{\alpha}_i^{n^-_{i}}$. \begin{lem} Let $T$ be a $d$ dimensional tiling which is composed of (possibly decorat ed) parallel epipeds and satisfies conditions {\rm D1,D2}. Then the homeomorphisms $\alpha_i$ defined above commute pairwise. Moreover, $\erz{\C}$ defined by (\ref{19061}) is regular. \end{lem} {\em Proof:} Draw in any tile\ which has \pe{d-1}s of type ${I_0\backslash\{i\}}$ a line segment joining the middle points of the boundary facets of this type. (The middle point of an \pe{n}\ of type $J$ is at $\sum_{i\in J}\frac{1}{2}\xi_i$.) This segment goes through the middle point of the tile. All the line segments in a tiling fit together to yield a set of infinite line s in $\mbox{\rm \twlset R}^d$. Such a line\ will be called $i$-line\ its type being $i$. As an example the below figure shows a patch of an octogonal tiling with its line segments of two types. The four vectors $\{\xi_1,\dots,\xi_4\}$ are given to the right, the line s are of type $1$ and $3$. \epsffile[0 0 420 260]{gFig1.ps} In general the line s have the properties: \begin{itemize} \item[1] If $l\neq l'$ then $l\cap l'$ contains at most one point (which is then the middle point of the tile through which they both go), and if $l$ and $l'$ are of the same type then $l\cap l'=\emptyset$. \item[2] Let $\pi_i$ be the orthogonal projection along the span of $\{\xi_j\}_{j\in I_0\backslash\{i\}}$. Then, for $d=2$ and any line\ which is of type $i$, $\pi_i(l)$ equals to the span of $\xi_i$. \end{itemize} Any type of line\ shall now be given the direction of the chosen normal of $I_0\backslash\{i\}$. The tile s belonging to it can be ordered and the application of $\breve{\alpha}_i$ can be geometrically interpreted as a shift to the next tile along an $i$-line . Condition D2 then reads that any line\ of type $i\in I_0$ contains in both directions infinitely many tile s of type $I_0$. Let us single out two elements of $I_0$ which we denote by $1$ and $2$. Let $l_0,m_0$ be the line s of type $1$ resp.\ $2$ which go through $x$, the pointed tile of $T_x\in\Omega_{\erz{\C}}$. Let $m_1,l_1,m'$ be the successor resp.\ of $m_0$ along $l_0$, of $l_0$ along $m_0$, and of $m_0$ along $l_1$, c.f.\ the below figure where this situation is indicated topologically. \epsffile[0 0 420 180]{gFig2.ps} Then $\alpha_1\circ\alpha_2(T_x)=\alpha_2\circ\alpha_1(T_x)$ if and only if $m'=m_1$. We first proof that this is the case for $d=2$. By property 1 above $m'$ can neither intersect $l_1$ at a second point nor $m_0$ at all nor can it intersect $l_0$ at a lower point than $m_1$ does. The same reasoning applies to $m_1$ so that, since $\pi_2(m')$ is the span of $\xi_2$, $m'\cap m\neq\emptyset$. Hence $m'=m_1$. The case $d>2$ can be traced back to $d=2$ reducing $T_x$ to a two dimensional tiling $\overline{T_x}$ by the following process. Let $J_0=I_0\backslash\{1,2\}$. We define the $J_0$-neighborhood of a tile $a$ which has \pe{d-2}s of type $J_0$ to consist of this tile an those tiles whose boundary intersects the boundary of $a$ at a \pe{d-2}\ of type $J_0$. Let $\overline{T_x}$ be the smallest connected component consisting of tiles having \pe{d-2}s of type $J_0$ in such a way that, first, it contains with a tile its $J_0$-neighborhood, and second, it contains the pointed tile of $T_x$. In particular $\overline{T_x}$ is a connected topological manifold which contains complete $1$- and $2$-line s. Let $\pi$ be the orthogonal projection along $\langle\{\xi_i\}_{i\in J_0}\rangle$. Then $\pi(\overline{T_x})$ is a two dimenional tiling to which we can apply the above arguement leading to $\pi(m_1)=\pi(m')$. To see that this implies $m_1=m'$ consider the homotopy $F_t(\xi_i)=(1-t)\xi_i+t\pi(\xi_i)$ for $i\notin J_0$ and $F_t(\xi_i)=\xi_i$ for $i\in J_0$. $F_1(\overline{T_x})$ may be understood as the Cartesian product of $\pi(\overline{T_x})$ with a $d\!-\!2$-facet of type $J_0$. Certainly no new intersections between lines in $F_t(\overline{T_x})$ occurr for $t>0$ so that the decomposition shows that $\pi(m_1)=\pi(m')$ whenever $m_1=m'$. By compactness of $\Omega_{\erz{\C}}$ the distance of two neighbored line s has an upper and a lower bound. Hence for some $r$ any $u\in{{\mTxx}^0}$ of radius larger than $r$ contains a tile of type $I_0$. Therefore $\erz{\C}$ is regular. \hfill$\Box$ \begin{thm} Let $T$ be a $d$ dimensional tiling which is composed of (possibly decorat ed) parallel epipeds and satisfies conditions {\rm D1,D2}. Then it reduces to a $Z^d$ decoration. \end{thm} {\em Proof:} The last lemma implies that the map $\Omega_{\erz{\C}}\times_\alpha \mbox{\rm \twlset Z}^d\to{\cal R}_{\Omega_{\erz{\C}}}$: $(\omega,e_i)\mapsto [\omega, c]$ where $c$ is uniquely determined by $c=c_1\dots c_{n^+(\omega)}$ with $c_k\in{\cal B}_i$, $c_k\vdash c_{k+1}$, and $L(c)\preceq\omega$, is an isomorphism of groupoids.\hfill$\Box$\bigskip All tilings which are obtained by the so-called generalized dual (or grid) method \cite{SoSt2} fall under the vicinity of the above theorem provided they satisfy D2. In fact, the lines of different type can be understood objects dual to the hyperplanes of the grids so that we just inverted the construction of \cite{SoSt2}. In particular all tilings obtained by the cut and projection method \cite{DuKa} (which is equivalent to the grid method using special grids only) reduce to decorations as they all satisfy D2. Among the substitution\ tilings which reduce to decorations of $\mbox{\rm \twlset Z}^2$ are the Ammann-Beenker- the Socolar- and the Penrose tilings. \section{$C^*$-algebra ic characterization of tilings} We review the $C^*$-algebra ic constructions neccessary to formulate the $K$-theoretical gap labelling for particles moving in a tiling. As already mentionned the $K$-theoretical gap labelling may be seen as an invariant of a $C^*$-algebra. The natural candidate for this algebra is the algebra $A_T$ associated to the tiling $T$ in which the particle is moving since it contains all local operators involving translations and multiplications with pattern dependent functions. \begin{df} The algebra ${\cal A}_T$ associated to a tiling $T$ is the reduced groupoid-$C^*$-algebra\ $C^*_{red}({\cal R})$ of the groupoid ${\cal R}$ associated to $T$. \end{df} Any $r$-dicrete groupoid defines a reduced groupoid-$C^*$-algebra\ \cite{Ren}. Specified to the groupoid ${\cal R}$ it is defined as follows: Let $C_c({\cal R})$ be the $*$-algebra of continuous functions $f:{\cal R}\rightarrow \mbox{\rm \twlset C}$ with compact support and multiplication and involution given by \begin{eqnarray} \label{25061} f*g\,[\omega,c] & = & \sum_{[\omega\cdot c,{c'}]\atop\omega,c\: \mbox{\tiny fixed}} f[\omega,cc']\: g[\omega\cdot cc',{c'}^{-1}] , \\ f^*\,[\omega,c] & = & \overline{f[\omega\cdot c,c^{-1}]} \label{25062}. \end{eqnarray} $C_c({\cal R})$ is generated by the characteristic functions $e_c$ onto ${\cal U}_c$ which satisfy the relations \begin{eqnarray} e_c*e_{c'} &=& \left\{ \begin{array}{ll} e_{c c'} & \mbox{if $c\vdash c'$} \\ 0 & \mbox{else} \end{array} \right.\\ e_c^*&=&e_{c^{-1}}. \end{eqnarray} Hence in any representation $e_c$ is represented as a partial isometry. $C^*_{red}({\cal R})$ is the closure of $C_c({\cal R})$ taken with respect to the reduced norm. This norm is defined using the family of representations $\pi_{[\omega]_o}$, one for each point in the non commutative\ space, acting on the Hilbert space of square summable functions $\Psi:[\omega]_o\rightarrow \mbox{\rm \twlset C}$ through \begin{equation} (\pi_{[\omega]_o}(e_{c})\Psi)(\omega)=\left\{ \begin{array}{ll} \Psi(\omega\cdot c) & \mbox{if $L(c)\preceq\omega$} \\ 0 & \mbox{else} \end{array} \right.. \end{equation} Now the reduced norm is given by $\|f\|_{red}=\sup_{[\omega]_o\in[\Omega]_o}\|\pi_{[\omega]_o}(f)\|$ where $\|\pi_{[\omega]_o}(f)\|$ is the operator norm. Since the topology of ${\cal R}$ has a countable basis ${\cal A}_T$ is separable.\bigskip As an example consider a tiling which is a decoration of $\mbox{\rm \twlset Z}^d$. That is there are $d$ commuting homeomorphisms $\alpha_i$ of the hull such that ${\cal R}=\Omega\times_\alpha\mbox{\rm \twlset Z}^d$. To any continuous $f:\Omega\times_\alpha\mbox{\rm \twlset Z}^d\to \mbox{\rm \twlset C}$ with compact support one may assign the function $\hat{f}:\mbox{\rm \twlset Z}^d\to C(\Omega)$ through $\hat{f}(k)(\omega) = f(\omega,k)$. Carried over from (\ref{25061},\ref{25062}) multiplication and involution then become convolution resp.\ involution twisted by $\alpha$: \begin{eqnarray} \hat{f}*\hat{g}\:(k)& =& \sum_{m\in\mbox{\rm \twlset Z}^d} \hat{f}(m)\, (\hat{g}(k\!-\!m)\circ\alpha(m)) \\ \hat{f}^*(k)& =& \overline{\hat{f}(-k)\circ\alpha(k)}. \end{eqnarray} The closure $C^*_{red}(\Omega\times_\alpha\mbox{\rm \twlset Z}^d)$ is isomorphic to $C(\Omega)\times_\alpha\mbox{\rm \twlset Z}^d$, the crossed product of $C(\Omega)$ with $\mbox{\rm \twlset Z}^d$ by the action $\alpha(k)(\hat{f}(m))=\hat{f}(m)\circ\alpha(k)$, $\alpha(k)=\alpha_1^{k_1}\circ\dots \circ \alpha_d^{k_d}$. \bigskip Remember that $T$ is minimal if the closure of the orbit of any $\omega$ equals $\Omega$. The lattice of closed (twosided) ideals of ${\cal A}_T$ may be identified with the lattice of open invariant subsets of $\Omega$, i.e.\ those open subsets which contain next to an element $T$ all its equivalent elements \cite{Ren}. The groupoid ${\cal R}$ is called minimal if $\Omega$ does not contain any proper invariant open subset. But this is just excluded for minimal tilings. So if $T$ is minimal then ${\cal R}$ is minimal which implies that ${\cal A}_T$ is simple \cite{Ren}. \bigskip Lemma~\ref{23051} has a $C^*$-algebra ic counterpart. Two $C^*$-algebra s ${\cal A}$ and ${\cal B}$ are called stably isomorphic if ${\cal A}\otimes{\cal K}$ is isomorphic to ${\cal B}\otimes{\cal K}$, where ${\cal K}$ is the algebra of compact operators (on an infinite dimensional separable Hilbert space). A theorem of Brown \cite{Bro} states that a full reduction ${\cal B}$ of ${\cal A}$ is stably isomorphic\ to ${\cal A}$. In this context a reduction of ${\cal A}$ is determined by a projection $p\in{\cal A}$, namely given by ${\cal A}_p:=\{x\in{\cal A}|px=xp=x\}$, and ${\cal A}_p$ is full if the twosided ideal generated by $p$ is dense in ${\cal A}$. \begin{lem} \label{22061} Let ${\cal N}\subset{\cal M}_{\rm I\!I}$ be regular. Then $C^*_{red}({\cal R}_{\Omega_{\cal N}})$ is stably isomorphic\ to ${\cal A}_T$. \end{lem} {\em Proof:} Let $\chi^{}_{\Omega_{\cal N}}$ be the projection in ${\cal A}_T$ which is the characteristic function onto $\Omega_{\cal N}$ and $f\in C_c({\cal R})$. Then $f*\chi^{}_{\Omega_{\cal N}}=\chi^{}_{\Omega_{\cal N}}*f=f$ if and only if $\mbox{supp}f\subset {\cal R}_{\Omega_{\cal N}}$. Hence $C_c({\cal R}_{\Omega_{\cal N}})$ is a dense subalgebra of $({\cal A}_T)_{\chi^{}_{\Omega_{\cal N}}}$ and, since the closure of $C_c({\cal R}_{\Omega_{\cal N}})$ with respect to the norm of ${\cal A}_T$ on the one hand and with respect to reduced norm defined for ${\cal R}_{\Omega_{\cal N}}$ on the other coincide, we have $({\cal A}_T)_{\chi^{}_{\Omega_{\cal N}}} = C^*_{red}({\cal R})_{\chi^{}_{\Omega_{\cal N}}}$. To show that this reduction is full we show that the two sided ideal generated by $\chi^{}_{\Omega_{\cal N}}$ contains $1$, the unit of the algebra, if ${\cal N}$ is regular. By Lemma~\ref{18071} there is a finite set $\{c_i\}_i$ satisfying $\Omega=\dot{\bigcup}_i U_{L(c_i)}$ and $U_{R(c_i)}\subset\Omega_{\cal N}$. Then $1=\sum_i e_{c_i}e_{c_i^{-1}}=\sum_i e_{c_i}\chi^{}_{\Omega_{\cal N}}e_{c_i^{-1}}$ showing that 1 is contained in the ideal.\hfill$\Box$ \subsection{$K$-theoretical gap labelling} A local selfadjoint operator $H$ describing the motion of a particle in the tiling $T$ is an operator in the representation $\pi_T$ of ${\cal A}_T$, namely \begin{equation} \label{14061} H\psi(T_x)=\sum_{x'\in X(T)} H_{x,x'}\psi(T_{x'}). \end{equation} Locality refers to the requirement that the matrix element $H_{x,x'}$ depends only on a the doubly pointed pattern class\ of a certain size to which $x,x'$ belong, i.e.\ that $H=\pi_T(h)$ for some $h\in C_c({\cal R})$. In particular $h$ may be of the form $h=-\Delta + \sum_i v_i e_{u_i}$ where $\{u_i\}_i$ is a collection of pointed pattern class es and $v_i\in\mbox{\rm \twlset R}$ and $\Delta$ is the discrete Laplacian. The latter takes the form $\Delta=\sum_{A\in{\mTxx}_{2,\neq}} b_A e_A -\sum_{a\in{\mTxx}_{1}} b_a e_a$ where $b_a\in\mbox{\rm \twlset R}$ and $b_A^*=b_{A^{-1}}\in\mbox{\rm \twlset C}$. The values of the integrated density of states $N_H(E)$ of $H$ at energies $E$ lying in a gap of its spectrum serve as labels for the gaps; they are insensitive to certain perturbations of the operator. In Bellissard's $K$-theoretic formulation of the gap labelling \cite{Be4,Be1} these values are recognized as elements of $\mbox{\rm tr}_*K_0({\cal A})$ where ${\cal A}$ is a $C^*$-algebra\ represented on the above Hilbert space, $H$ an element of its representation, $\mbox{\rm tr}$ a trace on ${\cal A}$, and $\mbox{\rm tr}_*$ the induced state on its $K_0$-group $K_0({\cal A})$: The $K_0$-group of a unital $C^*$-algebra\ ${\cal A}$ is obtained via Grothendieck's construction from the monoid of projection classes of ${\cal A}\otimes{\cal K}$, i.e.\ the equivalence classes of projections of ${\cal A}\otimes{\cal K}$ under $p\sim q$ whenever $\exists u\in{\cal A}\otimes{\cal K}: p = uu^*,q = u^*u$ may be added orthogonally, $[p]+[q]=[p\oplus q]$, yielding a monoid $V({\cal A})$, and this monoid may be completed to an abelian group whose elements are classes of pairs under $([p],[q])\sim ([p'],[q'])$ whenever $\exists [r]\in V(A):[p]+[q']+[r] = [q]+[p']+[r]$, see \cite{Bla,Mur} for details. Any trace of ${\cal A}$ extends to a trace on ${\cal A}\otimes{\cal K}$ and defines a linear map $\mbox{\rm tr}_*:K_0({\cal A})\to\mbox{\rm \twlset R}$: $[p]\mapsto \mbox{\rm tr}(p)$. The gap labelling in the above formulation requires the equality \begin{equation} \label{28071} N_H(E) = \mbox{\rm tr}(\chi_{h\leq E}) \end{equation} $h$ being the element which is represented by $H$ and $\chi_{h\leq E}$ the spectral projection of $h$ to energies smaller or equal to $E$. This equality involves validity of Shubin's formula \cite{Be1} by which the trace is equated to the operator trace per volume in the corresponding representation. Taking ${\cal A}$ to be the algebra associated to the tiling and considering the above representation the $K$-theoretical gap labelling then reads: if $E$ lies in a gap \begin{equation} N_H(E)\in \mbox{\rm tr}_*K_0({\cal A}_T)\cap[0,1] \end{equation} provided (\ref{28071}) holds. In other words $\mbox{\rm tr}_*K_0({\cal A}_T)\cap[0,1]$ is the set of gap labels predicted by $K$-theory. One motivation for using ${\cal A}_T$ is that it is expected not to yield to many values on the r.h.s.\ so that if the couplings $v_i$ are strong and diverse enough all elements of $\mbox{\rm tr}_*K_0({\cal A}_T)$ actually occur as labels for open gaps of $H$. In that case, and if $\mbox{\rm tr}$ is faithful on ${\cal A}_T$, the density of the values of the integrated density of states on gaps in $[0,1]$ expresses the fact that the continuous part of the spectrum is a Cantor set. In fact, conclusions on the nature of the spectrum may partly be drawn without the need to connect the gap labelling with the values of the integrated density of states, any faithful trace may be used. For instance if $\mbox{\rm tr}$ is faithful and the set of gap labels ${\cal L}_{\mbox{\rm tr}} (h)=\{\mbox{\rm tr}(\chi_{h\leq E})|E\notin\sigma(h)\}$ dense in $[0,1]$ the spectrum $\sigma(h)$ cannot contain a proper closed interval, for, if $[a,b]\in\sigma(h)$, then by faithfulness $\mbox{\rm tr}(\chi_{h\leq b})>\mbox{\rm tr}(\chi_{h\leq a})$ -- here $\mbox{\rm tr}$ has to be extended to measurable functions over $\sigma(h)$ -- so that $[0,1]\backslash{\cal L}_{\mbox{\rm tr}} (h)$ would contain the open interval $(\mbox{\rm tr}(\chi_{h\leq a}),\mbox{\rm tr}(\chi_{h\leq b}))$. However, up to now there is no $K$-theoretic formulation of a condition for a Schr\"odinger operator\ $h$ under which ${\cal L}_{\mbox{\rm tr}} (h)$ coincides with $\mbox{\rm tr}_*K_0({\cal A}_T)$. For a $C^*$-algebraic formulation of such a condition for the discrete magnetic Laplacian on $\mbox{\rm \twlset Z}^2$ see \cite{Sh2}. Another consequence of the realization of $H$ as an element of a representation of $A_T$ is that for minimal $T$ its spectrum has no discrete part, since the spectral projection onto the eigenspace of a discrete eigenvalue with finite multiplicity would have to be represented by a compact operator. In fact, for minimal $T$ the algebra ${\cal A}_T$ is antilimial \cite{Ped}, and hence does not contain compact operators, namely it is neither limial itself -- its primitive spectrum contains one single point whereas two different representations $\pi_{[\omega]_o}$ and $\pi_{[\omega']_o}$ are not unitarily equivalent -- nor can it contain a limial ideal. \subsubsection*{Traces on ${\cal A}_T$} Any normalized trace $\mbox{\rm tr}$ on ${\cal A}_T$ restricted to a linear functional $\mu$ on $C(\Omega)$ defines a normalized measure on $\Omega$ also denoted by $\mu$ through $\mu(f)=\mbox{\rm tr}(f)=\int f d\mu$. This measure is invariant under the groupoid in the sense that $\mu(U_{L(c)})=\mu(U_{R(c)})$. A direct consequence is that the values of $\mu$ on integer valued continuous functions over $\Omega$ lie in the $\mbox{\rm \twlset Z}$-module generated by the traces of projections of the algebra, and therefore \begin{equation} \label{11051} \mu(C(\Omega,\mbox{\rm \twlset Z})) \subset\mbox{\rm tr}_*K_0({\cal A}_T). \end{equation} The existence of a faithful trace implies that the notion of positive elements of $H({\cal R})={\cal C}(\Omega,\mbox{\rm \twlset Z})/E_{\cal R}$ as being those which have a representative which is a positive function turns $H({\cal R})$ into an ordered group. That is the positive elements $H^+({\cal R})$ satisfy $H^+({\cal R})+H^+({\cal R})\subset H^+({\cal R})$, $H^+({\cal R})-H^+({\cal R})=H({\cal R})$ and $H^+({\cal R})\cap -H^+({\cal R})=\{0\}$. In fact, the existence of a faithful trace rules out that a nonzero element can be both positive and negative. It being zero on $E_{\cal R}$ it moreover defines a state (if normalized) on $H({\cal R})$. A similar statement holds true for $K_0({\cal A}_T)$: The existence of a normalized faithful trace guarantees that ${\cal A}_T$ is stably finite and therefore the usual notion of positive elements of $K_0({\cal A}_T)$ as those which have a representative which is a projection in ${\cal A}_T\otimes{\cal K}$ turns $K_0({\cal A}_T)$ into an ordered group (and the trace induces a state on that group) \cite{Bla}. Conversely any ${\cal R}$-invariant normalized measure $\mu$ on the hull $\Omega$ defines a normalized trace through \begin{equation} \label{24061} \mbox{\rm tr}(f):=\int_\Omega\mbox{P}(f)\,d\mu \end{equation} where $\mbox{P}:\,C^*_{red}({\cal R})\rightarrow C(\Omega)$ is the restriction map. $\mbox{P}$ is the unique conditional expectation on $C(\Omega)$ and is faithful \cite{Ren}. Moreover if $T$ is minimal every non trivial invariant measure has to have closed support $\Omega$ so that $\mbox{\rm tr}$ defined by (\ref{24061}) is faithful.\bigskip One of the goals of this article is the determination of $\mbox{\rm tr}_*K_0({\cal A}_T)$. The question under which circumstances a given trace satisfies Shubin's formula and (\ref{28071}) will not be addressed here, but see \cite{Be1,BBG,Ke2} for investigations in this directions. \subsection{$K_0$-groups for tilings} The $K_0$-group of a $C^*$-algebra\ depends together with its order structure only on its stable isomorphism class. The same holds true for the $K_1$-group which may be understood as the $K_0$-group of the suspension of ${\cal A}$. For this reason we may apply the known results on crossed products to obtain the structure of the $K$-theory for tilings which reduce to decorations. We are not able to present any results on $K$-groups in case the tiling algebras are not stably isomorphic to crossed products.\bigskip A lot is known about the $K$-groups of crossed products of the form $C(\Omega)\times_\alpha\mbox{\rm \twlset Z}^d$ in particular for zero dimensional $\Omega$. Recently a relation of these $K$-groups with the group cohomology $H(\mbox{\rm \twlset Z}^d,C(\Omega,\mbox{\rm \twlset Z}))$ of $\mbox{\rm \twlset Z}^d$ with coefficients in $C(\Omega,\mbox{\rm \twlset Z})$ was discovered \cite{FoHu}. We will not discuss group cohomology here but what is important for us is that the cohomology group of highest nonvanishing degree (which is $d$) coincides with the group of coinvariants of $\Omega\times_\alpha\mbox{\rm \twlset Z}^d$ and that the $K_0$-group decomposes into \begin{equation} \label{08075} K_0(C(\Omega)\times_\alpha\mbox{\rm \twlset Z}^d)\cong H(\Omega\times_\alpha\mbox{\rm \twlset Z}^d)\oplus H' \end{equation} where $H'$ is trivial for $d=1$ and equal to $\mbox{\rm \twlset Z}$ in $d=2$. (In general $H'$ is a direct sum of cohomology groups of degrees $d-2n$, $0<n\leq \frac{d}{2}$ and $K_1(C(\Omega)\times_\alpha\mbox{\rm \twlset Z}^d)$ is a direct sum of cohomology groups of degrees $d-1-2n$, $0\leq n<\frac{d}{2}$ \cite{FoHu}.) Up to $d=3$ this result was obtained before \cite{Els2} in an even more explicit form in which in particular becomes clear that the image of the state on $K_0(C(\Omega)\times_\alpha\mbox{\rm \twlset Z}^d)$ induced by a trace on $C(\Omega)\times_\alpha\mbox{\rm \twlset Z}^d$ satisfies\footnote{ In \cite{Els2} ergodic measures have been used, but ergodicity not essential for the proof of (\ref{06061}).} \begin{equation} \label{06061} \mbox{\rm tr}_*K_0(C(\Omega)\times_\alpha\mbox{\rm \twlset Z}^d)=\mbox{\rm tr}_*(H(\Omega\times_\alpha\mbox{\rm \twlset Z}^d)) = \mu(C(\Omega,\mbox{\rm \twlset Z})). \end{equation} (\ref{06061}) holds for arbitrary $d$ under the restriction that $\Omega\times_\alpha\mbox{\rm \twlset Z}^d$ splits into a Cartesian product $\Omega\times_\alpha\mbox{\rm \twlset Z}^d=(\Omega_1\times_{\alpha_1}\mbox{\rm \twlset Z})\times \dots\times(\Omega_d\times_{\alpha_d}\mbox{\rm \twlset Z})$ \cite{Ke2}. (\ref{08075}) immediately carries over to tilings which reduce to decorations of $\mbox{\rm \twlset Z}^d$ due to Lemma~\ref{22061} \begin{equation} K({\cal A}_T)\cong K(C(\Omega_{\cal N})\times_\alpha\mbox{\rm \twlset Z}^d). \end{equation} This also concerns the order structure which is however not known for $d>1$.\medskip As for (\ref{06061}) recall that in case the tiling reduces to a decoration ${\cal A}_{\cal N}=C(\Omega_{\cal N})\times_\alpha\mbox{\rm \twlset Z}^d$ is a subalgebra of ${\cal A}_T$, i.e.\ there is an embedding $\imath:{\cal A}_{\cal N}\rightarrow{\cal A}_T$. This embedding induces an isomorphism $\imath_*$ from $K_0({\cal A}_{\cal N})$ onto $K_0({\cal A}_{T})$. In fact, for separable $C^*$-algebra s, ${\cal A}$ being stably isomorphic to ${\cal B}$ is equivalent to the existence of a (strong) Morita equivalence ${\cal A}$-${\cal B}$-bimodule which may be viewed as an element of $K\!K({\cal A},{\cal B})$ and is a special case of a $K\!K$-equivalence \cite{Bla,Ska,Cone}. Any $K\!K$-equivalence between ${\cal A}$ and ${\cal B}$ yields an isomophism from $K\!K(\mbox{\rm \twlset C},{\cal A})$ onto $K\!K(\mbox{\rm \twlset C},{\cal B})$, namely by multiplying it from the right, the multiplication being the Kasparov product. Translated into $K_0$-groups, $K\!K(\mbox{\rm \twlset C},{\cal A})$ being isomorphic to $K_0({\cal A})$, the right multiplication of elements of $K\!K(\mbox{\rm \twlset C},{\cal A}_{\cal N})$ with the canonical Morita equivalence ${\cal A}_{\cal N}$-${\cal A}_T$-bimodule, which as a linear space is ${\cal A}_{\cal N}{\cal A}_T$, precisely becomes $\imath_*$. Now $\mbox{\rm tr}\circ\imath$ is a trace on ${\cal A}_{\cal N}$ which is normalized to $\mbox{\rm tr}(\imath(1_{{\cal A}_{\cal N}}))=\mbox{\rm tr}(\chi^{}_{\Omega_{\cal N}})=\mu(\Omega_{\cal N})$. Since the invariant measure on $\Omega_{\cal N}$ corresponding to $\mbox{\rm tr}\circ\imath$ is $\mu|_{\Omega_{\cal N}}$ we get \begin{equation} \mbox{\rm tr}_*\circ \imath_*K_0({\cal A}_{\cal N})=\mu(C(\Omega_{\cal N},\mbox{\rm \twlset Z})) \end{equation} and therefore \begin{equation} \mbox{\rm tr}_*K_0({\cal A}_{T})=\mu(C(\Omega_{\cal N},\mbox{\rm \twlset Z}))\subset \mu(C(\Omega,\mbox{\rm \twlset Z})). \end{equation} Together with (\ref{11051}) this extends (\ref{06061}) to tilings which reduce to decorations of $\mbox{\rm \twlset Z}^d$, $d\leq 3$. $K_0({\cal A}_{\cal N})$ and $K_0({\cal A}_{T})$ differ only in their order units (the images of the units of the algebras in $K_0$). If one identifies them as above the order unit of the former is the $K_0$-class of $\chi^{}_{\Omega_{\cal N}}$. \section{Substitution tilings} We have seen that the integer group of coinvariants furnishes part of the $K$-theory of the algebra of the tiling. In particular in one and two dimensions it yields up to order the $K_0$-group. To determine this group we need further structure which is provided by a locally invertible substitution.\bigskip A substitution\ of a tiling may be thought of as a rule according to which the tile s of the tiling are to be replaced by pattern s which fit together to yield a new tiling. An algebraic way to formulate this is by means of homomorphisms of the almost-groupoid\ ${\cal M}_{\rm I\!I}$ incorporating thus their local nature. For the definition recall that $E^0(R(c))=\{e\in{\cal M}_{\rm I\!I}|R(e)=R(c)\}$ and that a homomorphism satisfies the growth condition with $t>0$ if $|\mbox{\rm rad}(\hat{\rho}(u))-t\,\mbox{\rm rad}(u)|$ is a bounded function on ${{\mTxx}^0}$. Let $x(c)$ be the first pointed tile of the doubly pointed pattern class\ $c$. \begin{df} A substitution\ of $T$ is a homomorphism of almost-groupoid s \begin{equation} \hat{\rho}:{\cal M}_{\rm I\!I}\to {\cal M}_{\rm I\!I} \end{equation} which satisfies the growth condition with $t>1$, and for which the pattern class\ of $\hat{\rho}(c)$ is composed without overlap (up to boundaries) of the pattern class es of the $\hat{\rho}(x(e))$, $e\in E^0(R(c))$, and the first resp.\ second pointed tile of $\hat{\rho}(c)$ correspond to the pointed tile of $\hat{\rho}(x(c))$ resp.\ $\hat{\rho}(x(c^{-1}))$. \end{df} Since $\hat{\rho}$ satisfies the growth condition it defines a homomorphism of ${\cal R}$ into itself. We shall call a tiling resp.\ its class invariant under a substitution\ if $[\rho]_o(T)=T$. We call it a substitution\ tiling if it allows for a substitution\ such that $[\rho]_o(T)$ is locally isomorphic to $T$. Like any homomorphism $\hat{\rho}$ is determined by its action on the generators ${\mTxx}_{2,\neq}$. But by the above definition this means that $\hat{\rho}$ may be given, first, by its image on the pointed pattern class es of tiles, ${\mTxx}_{1}$, and second, by the relative position of these images. A substitution\ can be iterated and it is called primitive if for some $n$ and all $a\in {\mTxx}_{1}$ the pattern class\ $\rho^n(a)$ contains at least one tile of each tile class.\medskip If a tiling allows for a substitution\ which has a right inverse, and this is equivalent with locally invertible as we have seen, then the substitution\ can be used to compute the coinvariants associated to the tiling class. We shall carry this out below but first give a brief discussion on geometric realizations of substitution s. \subsection{Deflation} Geometrically we view a pointed tiling as a representative of $T$ in the Euclidian space $\mbox{\rm \twlset R}^d$ with fixed origin $0$. To be precise we choose a point for each tile class (its puncture). Then $T_x$ shall correspond to the representative of $T$ for which the puncture of the pointed tile coincides with $0$ thereby getting a bijective correspondance between pointed tilings and representatives in $\mbox{\rm \twlset R}^d$ having the puncture of one of its tiles at $0$. Substitutions appear then as deflations followed by rescaling. Let $t^{-1}T$ denote $T$ rescaled by $t^{-1}$ and ${\cal M}(T)$ the (unpointed) pattern class es of $T$. A deflation is given by: \begin{itemize}\item[1] a pattern class $\rho_t(a)\in{\cal M}(t^{-1}T)$ for each tile class $a$, $t>0$. \item[2] a relative position between $a$ and $\rho_t(a)$ for each tile class $a$. \end{itemize} Let ${\cal T}$ be a representative of $T$ in $\mbox{\rm \twlset R}^d$. In the process of applying the deflation any tile of class $a$ in ${\cal T}$ is to be replaced by a pattern of class $\rho_t(a)$ (its replacement) at the relative position given above. To be more precise choose a tile in $\rho_t(a)$. Its puncture shall now indicate the position of a pattern of class $\rho_t(a)$ in $\mbox{\rm \twlset R}^d$. Then the relative position i.e.\ the difference between the position of the tile and its replacement shall be $x_a\in\mbox{\rm \twlset R}^d$. Thus if a tile of ${\cal T}$ of class $a$ is at $x\in\mbox{\rm \twlset R}$ then a pattern of class $\rho_t(a)$ is in $\rho_t({\cal T})$ at $x-x_a$. The object $\rho_t({\cal T})$ is a composition of patterns of the classes $\rho_t(a)$, and it is required that this yields a tiling which is locally isomorphic to $t^{-1}T$. In particular no overlap and no gaps are allowed. Such a deflation defines a substitution\ in the algebraic sense (as a homomorphism of ${\cal M}_{\rm I\!I}$). By the above choice of a tile in $\rho_t(a)$ the latter may now be understood as a pointed pattern class\ of $t^{-1}T$. We let $\hat{\rho}(a)$ be $t\rho_t(a)$ which is $\rho_t(a)$ scaled by $t$, the pattern class\ of $\hat{\rho}(M_{x_1x_2})$ be $t\rho_t(M)$, and the first resp.\ second pointed tile be the one corresponding to the pointed tile of $\hat{\rho}(x(M_{x_1x_2}))$ resp.\ $\hat{\rho}(x(M_{x_2x_1}))$.\medskip A deflation is locally invertible if the determination of whether or not a pattern\ of $\rho_t({\cal T})$ is a replacement of a tile\ of ${\cal T}$ may be uniquely carried out by inspection of the larger patterns around that pattern\ up to a given finite size. In other words, for each $a$ there is a finite set of pointed pattern classes $\Phi_t(a)\subset\mtx{t^{-1}T}$ with $\forall v\in\Phi_t(a):\rho_t(a)\preceq v$ such that whenever one of them occurrs in $\rho_t({\cal T})$ at $x\in\mbox{\rm \twlset R}$ then a tile of class $a$ occurrs in ${\cal T}$ at $x+x_a$. This condition allows one to locally obtain a preimage of any representative of any $t^{-1}T'$, $T'\in[\Omega]_o$. But this furnishes also a right inverse of $\rho$ on $\Omega$ as follows: Let ${\cal T}$ be the representative corresponding to $\omega\in\mbox{\rm im}\,\rho$. Then a preimage of $\omega$ under $\rho$ is given by the pointed tiling corresponding to $\rho_t^{-1}(t^{-1}{\cal T})-x_a$ where $t^{-1}{\cal T}$ is rescaled in such a way that the puncture remains on $0$ and $a$ is determined through the unique replacement $\rho_t(a)$ whose puncture is on $0$. Clearly the right inverse of $\rho$ on $\Omega$ extends to ${\cal S}$. And since non periodicity is forced by the existence of a locally invertible deflation \cite{GrSh,Ke2} the corresponding substitution\ $\hat{\rho}$ is locally invertible in the sense of section~\ref{11071}. \subsection{Path spaces over graphs and their dimension groups} This section is meant to fix the notation thereby giving an overview on the structures that will be needed.\bigskip A graph $\Sigma$ is a set of vertices $\Sigma^{(0)}$ and a set of edges $\Sigma^{(1)}$ with two maps $s,r:\Sigma^{(1)}\to \Sigma^{(0)}$, the source and the range map. Its connectivity matrix\ is the $|\Sigma^{(0)}|\times|\Sigma^{(0)}|$ matrix with coefficients \begin{equation} \sigma_{xy} := \mbox{number of edges which have source $y$ and range $x$}. \end{equation} A path $\xi=\xi_1\dots\xi_n$ of length $n$ over $\Sigma$ is a sequence of $n$ edges such that $r(\xi_k)=s(\xi_{k+1})$. We denote its lentgh by $|\xi|$. One sets $r(\xi)=r(\xi_{n})$ and $s(\xi)=s(\xi_{1})$. Two paths $\xi,\xi'$ with $r(\xi)=s(\xi')$ may be concatenated to yield longer path $\xi\circ\xi'=\xi_1\dots\xi_{|\xi|}\xi'_1\dots\xi'_{|\xi'|}$. The set of all (half) infinite paths over the graph yields a space which carries a compact metric topology, it is called the path space ${\cal P}_{\Sigma}$ of the graph. Its topology is generated by sets \begin{equation} U_{\xi} = \{\xi\circ\gamma|s(\gamma)=r(\xi)\}. \end{equation} These sets are closed as well so that ${\cal P}_{\Sigma}$ is a zero dimensional. Two (infinite) paths $\gamma$, $\gamma'$ are called confinal if for some $n$, $\forall i\geq n: \gamma_i=\gamma'_i$. Cofinality is an equivalence relation and the subspace of ${\cal P}_{\Sigma}\times{\cal P}_{\Sigma}$ of all cofinal pairs may be given a topology which is generated by \begin{equation} {\cal U}_{\xi\xi'} = \{(\xi\circ\gamma,\xi'\circ\gamma)|s(\gamma)=r(\xi)\} \end{equation} where it is required that $|\xi|=|\xi'|$ and $r(\xi)=r(\xi')$. With this topology the groupoid defined by the equivalence relation becomes an $r$-discrete principal groupoid ${\cal R}_\Sigma$. Its groupoid-$C^*$-algebra\ ${\cal A}_{\Sigma}$ is finitely approximated, i.e.\ an $AF$-algebra. The integer group of coinvariants of $H({\cal R}_\Sigma)$ of this groupoid is $C({\cal P}_{\Sigma},\mbox{\rm \twlset Z})/E_\Sigma$ where $E_\Sigma$ is generated by elements of the form $\eta({\cal U}_{\xi\xi'}) = \chi^{}_{U_\xi}-\chi^{}_{U_{\xi'}}$. $C({\cal P}_{\Sigma},\mbox{\rm \twlset Z})/E_\Sigma$ carries an order structure given by the notion of positive functions, e.g.\ $[f]_\Sigma>0$ whenever it has a representative $f>0$. With that order structure and the constant function $1$ representing the order unit $H({\cal R}_\Sigma)$ is also the dimension group or scaled ordered $K_0$-group of the $AF$-algebra ${\cal A}_{\Sigma}$ \cite{Eff}. We will not discuss the $C^*$-algebra ic and $K$-theoretic details here but only concentrate on the computation of the dimension group which will be used lateron to determine the coinvariants of ${\cal R}$. The main point is that $H({\cal R}_\Sigma)$ can be obtained by taking the algebraic limit of the directed system $(G_n,\sigma)$ where $G_n=\mbox{\rm \twlset Z}^{|\Sigma^{(0)}|}$ and $\sigma:G_n\to G_{n+1}$ is the homomorphism given by the connectivity matrix\ once the vertices have been identified with the standard base. The algebraic limit of the above system is a universal object which is a group $G$ together with homomorphisms $j_n:G_n\to G$ such that $j_{n+1}\circ \sigma = j_n$ \cite{Lan}. It is up to isomorphism determined by the property that, if there is any other group $G'$ and $j'_n:G_n\to G'$ such that $j'_{n+1}\circ \sigma = j'_n$ then there is a unique homomorphism $j:G\to G'$ such that $j'_n=j\circ j_n$. Moreover, $G$ inherits an order structure and the order unit from the standard order structures on the $G_n$. And $H({\cal R}_\Sigma)$ coincides with $G$ as ordered group with order unit. There are several "standard" realizations for this limit \cite{Lan,Mur}. We will use neither of them here but instead one which is less general but more suitable for our means, c.f.\ \cite{Hos}. In this realization the group is isomorphic to the quotient \begin{equation}\label{17061} H({\cal R}_\Sigma) \cong \{x\in \mbox{\rm \twlset R}^k|\exists n\geq 0:\sigma^n(x)\in\mbox{\rm \twlset Z}^k\}/ \{x\in \mbox{\rm \twlset R}^k|\exists n\geq 0:\sigma^n(x)=0\} \end{equation} with $k=|\Sigma^{(0)}|$. The maps $j_{n}:G_{n}=\mbox{\rm \twlset Z}^k\to H({\cal R}_\Sigma)$ are given by $j_{n}=\pi\circ \sigma^{-n}$ where $\sigma^{-1}$ is taking the preimage. If $G'$ and $j'_n:G_n\to G'$ with $j'_{n+1}\circ \sigma = j'_n$ is any other realization then $j(\pi(x))=j'_n(\sigma^n(x))$ for $\sigma^n(x)\in\mbox{\rm \twlset Z}^k$ yields the unique homomorphism $j:H({\cal R}_\Sigma)\to G'$ satisfying $j'_n=j\circ j_n$. To express the order structure in this realization we restrict for simplicity to the case where $\sigma$ is primitive so that it has a Perron-Frobenius-eigen\-value\ $\tau$ with left-Perron-Frobenius-ei\-gen\-vec\-tor\ $\nu$. If $\pi$ denotes the canonical projection of the above quotient, the positive cone is \begin{equation} H^+({\cal R}_\Sigma)\cong\{\pi(x)\in H({\cal R}_\Sigma)|\sum_i\nu_i x_i >0\}\cup\{0\}, \end{equation} and the order unit $\pi(w)\in H({\cal R}_\Sigma)$ with $w_i=1$. Since $\sigma$ is primitive, there is a unique normalized measure $\mu$ on ${\cal P}_{\Sigma}$ which is invariant under the groupoid ${\cal R}_\Sigma$, i.e.\ satisfies $\mu(U_\xi)=\mu(U_{\xi'})$ in case $|\xi|=|\xi'|$ and $r(\xi)=r(\xi')$ \cite{Eff}. If the left Perron-Frobenius-ei\-gen\-vec\-tor\ of $\sigma$ normalized to $\sum_i\nu_i=1$ then $\mu(U_\xi)=\tau^{-|\xi|}\nu_{r(\xi)}$ and its range on $C({\cal P}_{\Sigma},\mbox{\rm \twlset Z})$ is given by \begin{equation}\label{27071} \mu(C({\cal P}_{\Sigma},\mbox{\rm \twlset Z})) = \{\tau^{-n}\sum_i\nu_i n_i|n\geq 0,n_i\in\mbox{\rm \twlset Z}\}. \end{equation} In particular $\mu$ is well defined on $H({\cal R}_\Sigma)$ and the order may be expressed as \begin{equation} \label{11073} H^+({\cal R}_\Sigma) = \{x\in H({\cal R}_\Sigma)|\mu(x)>0\}\cup\{0\} . \end{equation} Elements which are neither positive nor negative are called infinitesimal. \subsection{Path spaces determined by substitution s} Any substitution, invertible or not, defines an obvious graph which however does not even for all locally invertible substitution s contain enough information of the tiling. In analogy to the one dimensional case we call it with \cite{For} the improper graph. It has the substitution\ matrix as connectivity matrix. But other graphs may also be attached to the substitution. These are related to the improper graph and coincide with it in the border forcing case. They allow for a coding of the tilings which yields a homeomorphism between their path space and the hull. This will be used to solve the $K$-theoretical gap labelling for a certain class of substitution\ tilings. \bigskip Let for $a\in{\mTxx}_{1}$ \begin{equation} E^n(a) = \{e\in{\cal M}_{\rm I\!I}|R(e)=\hat{\rho}^n(a)\}. \end{equation} The improper graph $\Sigma$ is the graph which has vertices and edges resp.\ \begin{eqnarray} \Sigma^{(0)}&=&{\mTxx}_{1} \\ \Sigma^{(1)}&=&\bigcup_{a\in{\mTxx}_{1}}\{(e,a)|e\in E^1(a)\} \end{eqnarray} and range resp.\ source maps $r,s:\Sigma^{(1)}\to\Sigma^{(0)}$ given by \begin{eqnarray} r(e,a)&=&a \\ s(e,a)&=&x(e). \end{eqnarray} Thus the edges of the graph which have range $a$ are in bijective correspondence to the tiles in $\hat{\rho}(a)$. $\Sigma$ has as connectivity matrix\ \begin{equation} \sigma_{a_1a_2} = |\{e\in E^1(a_1)|x(e)=a_2\}|, \end{equation} i.e.\ $\sigma_{a_1a_2}$ equals the number of $a_2$'s in $\hat{\rho}(a_1)$. This matrix (or sometimes its transpose) is also called the substitution\ matrix of $\hat{\rho}$. A path of finite length like $((e_1,a_1),\dots,(e_n,a_n))$ depends due to the particular form of the range map only on $(e_1,\dots,e_n)$ and $a_n$. It shall be written shorter $(e_1,\dots,e_n;a_n)$. \bigskip The improper graph does not contain enough information of the tiling to yield the right integer group of coinvariants. To improve this we have to incorporate the neighborhood of substitute s. To a given doubly pointed pattern class\ $c$ let ${{\cal F}}(c)$ be the set of all possible neighborhoods of $c$, i.e.\ the set of pattern class es occuring in $T$ which are composed of $c$ together with all the tiles the boundaries of which have a non empty intersection of with $c$. The pointed tiles are those of $c$. An integer $\upsilon\geq 0$ will parametrize a set of graphs $\Lambda_\mb$. (This generality is needed lateron for applications.) Define \begin{equation} \label{30061} {\cal B}^\upsilon_1:=\{(a,f)|a\in{\mTxx}_{1}, f\in{{\cal F}}(\hat{\rho}^\upsilon(a)),\rho^\upsilon(U_a)\cap U_f\neq\emptyset\}. \end{equation} The extra condition insures that only those neighborhoods of $\hat{\rho}^\upsilon(a)$ are taken into account which are neighborhoods of the $\upsilon$-fold substitute\ of $a$. Let $\Lambda_\mb$ be the graph with vertices and edges \begin{eqnarray} \Lambda_\mb^{(0)}&=&{\cal B}^\upsilon_1\\ \Lambda_\mb^{(1)}&=&\{(e,a,f)|(a,f)\in{\cal B}^\upsilon_1,e\in E^1(a)\} \end{eqnarray} and range and source map given by \begin{eqnarray} r(e,a,f)&=&(a,f)\\ s(e,a,f) &=& (x(e),f') \mbox{ with $f'\preceq L(\hat{\rho}^\upsilon(e)\hat{\rho}(f))$}. \end{eqnarray} (The $f'\in{\cal F}(\hat{\rho}^\upsilon(a))$ for which $f'\preceq L(\hat{\rho}^\upsilon(e)\hat{\rho}(f))$ is unique.) The connectivity matrix\ $\lambda_\mb$ of $\Lambda_\mb$ has coefficients \begin{equation} {\lambda_\mb\,}_{(a_1,f_1)(a_2,f_2)} = |\{e\in E^1(a_1)|x(e)=a_2,f_2\preceq L(\hat{\rho}^\upsilon(e)\hat{\rho}(f_1))\}|. \end{equation} Like for $\Sigma$ a path $((e_1,a_1,f_1),\dots,(e_n,a_n,f_n))$ depends only on $(e_1,\dots,e_n)$ and $(a_n,f_n)$ and may be abbreviated as $(e_1,\dots,e_n;a_n,f_n)$. Note that such a path satisfies $R(e_k)=\hat{\rho}(x(e_{k+1}))$ and $R(e_n)=\hat{\rho}(a_n)$. Therefore the map $\delta:{\cal P}_{\Lambda_\mb}^n(a,f)\to E^n(a)$ given by \begin{equation} \label{16062} \delta(e_1,\dots,e_n;a,f) := e_1\hat{\rho}(e_2)\dots\hat{\rho}^{n-1}(e_n) \end{equation} is a bijection for fixed $(a,f)$. Furthermore let $\beta_\upsilon:{\cal B}^\upsilon_1\to {\mTxx}_{1}$: \begin{equation} \beta_\upsilon(a,f) = a. \end{equation} Extending $\beta_\upsilon$ first to edges via $\beta_\upsilon(e,a,f):=(e,a)$ and then to paths on obtains a map $\beta_\upsilon:{\cal P}_{\Lambda_\mb}^n\to{\cal P}_{\Sigma}^n$: \begin{equation} \beta_\upsilon(e_1,\dots,e_n;a_n,f_n)=(e_1,\dots,e_n;a_n) \end{equation} which is not only surjective but also injective on ${\cal P}_{\Lambda_\mb}^n(a,f)$ for fixed $f$. In particular $\beta_\upsilon$ extends to a continuous surjective map from ${\cal P}_{\Lambda_\mb}$ onto ${\cal P}_{\Sigma}$. \begin{lem} \label{18061} Let $T$ be a substitution\ tiling with locally invertible substitution\ $\hat{\rho}$. For each $\omega\in\Omega$ there is a unique $(e,a)\in\Sigma^{(1)}$ such that $\omega\cdot e\in\mbox{\rm im}\,\rho$. \end{lem} {\em Proof:} By definition of a substitution\ does for any $c\in{\cal M}_{\rm I\!I}$ hold the alternative: either $c=c^{-1}$ or $\forall e_i\in E^1(x_i(c)):\:e_1\rho(c) e_2^{-1} \neq (e_1\rho(c) e_2^{-1})^{-1}$ where $x_1(c)=x(c)$ and $x_2(c)=x(c^{-1})$. Recall from the proof of Theorem~\ref{25071} that $\rho(\Omega)=\Omega_{\cal N}$ where ${\cal N}=\Phi({\cal M}_{\rm I\!I})$. Since for all $d\in\Phi(c)$, $\hat{\rho}(c)$ and $d$ have a common greater element the above alternative carries over for any $d\in {\cal N}$ in the form: either $d=d^{-1}$ or $\forall e_i\in E^1(x_i(d)):\:e_1 d e_2^{-1} \neq (e_1 d e_2^{-1})^{-1}$. Since $[\rho]_o(T)$ is locally isomorphic to $T$, ${\cal N}$ has to be regular and there is a $c\in{\cal M}_{\rm I\!I}$ such that $\omega\cdot c\in \Omega_{{\cal N}}$. And as ${\cal N}$ is approximating and generating ($T$ has to be non periodic) $c$ can be written as $c=e d$ where $d\in{\cal N}$ and $e\in E^1(x(d))$. We claim that $e$ is unique, for if not then $L(c)=e d{d'}^{-1}e'^{-1}$ and since this is a unit $d{d'}^{-1}$ must be equal to its inverse by the above alternative. Hence $e=e'$. Finally, if $\omega\cdot ed\in\mbox{\rm im}\,\rho$ then also $\omega\cdot e\in\mbox{\rm im}\,\rho$. The condition $a\preceq \rho^{-1}(\omega\cdot e)$ determines uniquely the $a$ for which $e\in E^1(a)$. \hfill$\Box$\bigskip Let ${\kappa_\mb}:\Omega\to\Lambda_\mb^{(1)}$: $\omega\mapsto(e,a,f)$ where $(e,a)$ is determined by the above lemma and $f\preceq\rho^{\upsilon-1}(\omega\cdot e)$. Furthermore, the first component of $(e,a)$ shall be used to define an extension of $\rho^{-1}$ to all of $\Omega$ through \begin{equation} \psi(\omega):=\rho^{-1}(\omega\cdot e). \end{equation} \begin{lem} $s({\kappa_\mb}(\psi^{n}(\omega)))=r({\kappa_\mb}(\psi^{n-1}(\omega))$. \end{lem} {\em Proof:} Let ${\kappa_\mb}(\psi^{n}(\omega))=(e_2,a_2,f_2)$ and ${\kappa_\mb}(\psi^{n-1}(\omega))=(e_1,a_1,f_1)$. We have to show that, first, $x(e_2)=a_1$, and second, that the $f'\in{\cal F}(\hat{\rho}^\upsilon(a_2))$ which satisfies $f'\preceq L(\hat{\rho}^\upsilon(e_2)\hat{\rho}(f_2))$ equals to $f_1$. The first follows from the definitions: $a_1\preceq\rho^{-1}(\psi^{n-1}(\omega)\cdot e_1)=\psi^n(\omega)$ and $L(e_2)\preceq \psi^n(\omega)$. As for the second, $f_2\preceq\rho^{\upsilon-1}(\psi^n(\omega)\cdot e_2)$ implies that $\rho(L(\hat{\rho}^{\upsilon-1}(e_2)f_2))\preceq\rho^\upsilon(\psi^n(\omega))= \rho^{\upsilon-1}(\psi^{n-1}(\omega)\cdot e_1)$. Hence $f'\preceq\rho^{\upsilon-1}(\psi^{n-1}(\omega)\cdot e_1)$ which is the relation determining $f_1$.\hfill$\Box$\bigskip As a consequence we may define a coding which is a map $Q_\mb:\Omega\to{\cal P}_{\Lambda_\mb}$ where the $n$th edge of $Q_\mb(\omega)$ is \begin{equation} Q_\mb(\omega)_n:= {\kappa_\mb}(\psi^{n-1}(\omega)). \end{equation} \begin{thm} \label{16063} Under the conditions of Lemma~\ref{18061} is $Q_\mb:\Omega\to{\cal P}_{\Lambda_\mb}$ a homeomorphism. \end{thm} {\em Proof:} Local invertibility implies that, for $n\geq\upsilon$ \begin{equation} \label{23062} {\cal V}_\upsilon^n(a,f,e_1,e_2) := \{[\omega\cdot e_1^{-1},e_1\hat{\rho}^{n-\upsilon}(f) e_2^{-1}]| \omega \in \rho^n(U_a)\cap\rho^{n-\upsilon}(U_f)\} \end{equation} is compact and open for all $(a,f)\in{\cal B}^\upsilon_1$ and $e_i\in E^n(a)$. (In fact, these are ${\cal R}$-sets.) Let $\xi$ be a path of length $n$ with $r(\xi)=(a,f)$. $Q_\mb(\omega)\in U_\xi$ if and only if $L(\delta(\xi))\preceq\omega$, $a\in\psi^n(\omega)$, and $f\preceq \rho^\upsilon(\psi^{n}(\omega))$. Since $\omega\cdot \delta(\xi) \in \mbox{\rm im}\,\rho^n$ and $\psi^n(\omega)=\rho^{-n}(\omega\cdot\delta(\xi))$ this means \begin{equation} \label{18062} Q_\mb^{-1}(U_\xi)={\cal V}_\upsilon^{|\xi|}(r(\xi),\delta(\xi),\delta(\xi)). \end{equation} Hence $Q_\mb$ is continuous. Since $Q_\mb^{-1}(U_\xi)$ is not empty, any path over $\Lambda_\mb$ is an accumulation point of $\mbox{\rm im}\, Q_\mb$. As the latter is closed $Q_\mb$ is surjective. Finally, to prove injectivity, let $r(Q_\mb(\omega)_k)=(a_k,f_k)$ and $\tilde{e}_n:=\delta(Q_\mb(\omega)_1\dots Q_\mb(\omega)_n)$. Then $\hat{\rho}^{n-\upsilon}(f)\preceq\omega\cdot\tilde{e}_n$ and $$u_n(\omega):=L(\tilde{e}_{n}\hat{\rho}^{n-\upsilon}(f_n))\preceq \omega.$$ Since $L(\tilde{e}_{n}\hat{\rho}^{n}(a_n))\preceq L(\tilde{e}_{n}\hat{\rho}^{n-\upsilon}(f_n))$ the radius of $u_n(\omega)$ diverges exponentionally with $n$. Hence $u_n(\omega)$ is an approximation which approximates $\omega$ and is uniquely determined by $Q_\mb(\omega)$. Thus $Q_\mb$ is injective. \hfill$\Box$\bigskip Theorem~\ref{16063} implies that $Q_\mb^*:C({\cal P}_{\Lambda_\mb},\mbox{\rm \twlset Z})\to C(\Omega,\mbox{\rm \twlset Z})$: $Q_\mb^*(f)=f\circ Q_\mb^{-1}$ is an isomorphism of groups. We may extend $Q_\mb$ to a map $Q_\mb:{\cal S}\to{\cal S}_{\Lambda_\mb}$ by restricting $Q_\mb\timesQ_\mb$ to ${\cal S}$ and since $T$ is not periodic we may view it as a map $Q_\mb:{\cal R}\to{\cal R}_{\Lambda_\mb}$. Then alike (\ref{18062}) \begin{equation} Q_\mb^{-1}({\cal U}_{\xi\xi'}) = {\cal V}^{|\xi|}_\upsilon(r(\xi),e_\xi,e_{\xi'}) \label{08071} \end{equation} showing that this map is continuous and surjective, but it is not injective. Since $\eta\circQ_\mb^{-1}=Q_\mb^*\circ\eta$ the image under $Q_\mb^*$ of the subgroup $E_{\Lambda_\mb}$ is contained in $E_{\cal R}$. Hence there is an induced surjective homomorphism $[Q_\mb^*]:H({\cal R}_{\Lambda_\mb})\to H({\cal R})$. Its kernel is $E_{\cal R}/Q_\mb^*(E_{\Lambda_\mb})$. In particular, since $\lambda_\mb$ is primitive, we get a corrolary from Theorem~\ref{16063}: \begin{cor} Let $T$ be a substitution\ tiling with primitive locally invertible substitution\ $\hat{\rho}$ which reduces to a decoration of $\mbox{\rm \twlset Z}^d$, $d\leq 3$, and $\mbox{\rm tr}$ be a trace on $A_T$. Then \begin{equation} \mbox{\rm tr}_*(K_0({\cal A}_T)) = \{\tau^{-n}\sum_i\nu_i n_i|n>0,n_i\in\mbox{\rm \twlset Z}\} \end{equation} where $\tau$ is the Perron-Frobenius-eigen\-value\ and $\nu$ the left Perron-Frobenius-ei\-gen\-vec\-tor\ of $\lambda_\mb$ normalized to $\sum_i\nu_i=1$. \end{cor} {\em Proof:} If $\mu$ is the measure on $\Omega$ obtained by restricting the trace then $\mu\circQ_\mb^{-1}$ is a measure on ${\cal P}_{\Lambda_\mb}$ which is invariant under ${\cal R}_{\Lambda_\mb}$. Using (\ref{06061}) one obtains $\mbox{\rm tr}_*(K_0({\cal A}_T)) = \mu(Q_\mb^{-1}(C({\cal P}_{\Lambda_\mb},\mbox{\rm \twlset Z})))$ and with (\ref{27071}) the statement.\hfill$\Box$\bigskip This solves the $K$-theoretical version of the gap labelling for substitution\ tilings with primitive locally invertible substitution. \bigskip {\em Remark:} Let us only remark here that one obtains not only the above embedding of groups by $Q_\mb^*$ but in fact a unital embedding $i:{\cal A}_{\Lambda_\mb}\rightarrow{\cal A}_T$ of the $AF$-algebra defined by $\Lambda_\mb$ into the algebra associated to the tiling similar to embedding of ${\cal A}_{\Sigma}$ in ${\cal A}_T$ described in \cite{Ke2}. In fact, a glance on (\ref{08071}) tells us that the characteristic functions on ${\cal V}^{|\xi|}_\upsilon(r(\xi),e_\xi,e_{\xi'})$ generate ${\cal A}_{\Lambda_\mb}$ as a subalgebra of ${\cal A}_T$. The embedding induces an order homomorphism $i_*$ of $K$-groups. Then $K_0({\cal A}_{\Lambda_\mb})=H({\cal R}_{\Lambda_\mb})$ as ordered group with order unit, and, under the hypothesis (\ref{08075}), $i_*:K_0({\cal A}_{\Lambda_\mb})\to H({\cal R})$ coincides with $[Q_\mb^*]$. However, in case $i_*$ is not surjective, i.e.\ the dimension of the tiling is bigger than one, our analysis will not determine the order structure on $K_0({\cal A}_T)$ but only a subcone of the positive cone $K^+_0({\cal A}_T)$. \subsection{The integer group of coinvariants for substitution\ tilings} Recall that the integer group of coinvariants is given by $H({\cal R})=H({\cal R}_{\Lambda_\mb})/\ker [Q_\mb^*]$. To compute $\ker [Q_\mb^*]=E_{\cal R}/Q_\mb^*(E_{\Lambda_\mb})$ we need to control pattern class es consisting of two tiles. In analogy to (\ref{30061},\ref{23062}) define\footnote{In principle the $\upsilon$ below could be chosen different from the one above, but we shall not make use of this generality.} for $n\geq\upsilon$ \begin{equation} {\cal B}^\upsilon_{2\neq}:= \{(A,F)|A\in{\mTxx}_{2,\neq},F\in{\cal F}(\hat{\rho}^\upsilon(A)),\rho^\upsilon(U_A)\cap U_F\neq\emptyset\}. \end{equation} \begin{equation} \label{28061} {\cal V}_\upsilon^n(A,F,e_1,e_2) := \{[\omega\cdot e_1^{-1},e_1\hat{\rho}^{n-\upsilon}(F) e_2^{-1}]| [\omega,\hat{\rho}^{n-\upsilon}(F)] \in \rho^n(U_A)\cap\rho^{n-\upsilon}(U_F)\} \end{equation} where $e_i\in E^{n}(x_i(A))$, $x_1(c)=x(c)$ and $x_2(c)=x(c^{-1})$. If $e_i$ are units we write ${\cal V}_\upsilon^n(A,F)$ and resp.\ ${\cal V}_\upsilon^n(a,f)$ for the above. It should be clear that for fixed $e_1,e_2$ the sets ${\cal V}_\upsilon^n(x,y,e_1,e_2)$ are for different $(x,y)\in{\cal B}^\upsilon_1\cup{\cal B}^\upsilon_{2\neq}$ pairwise disjoint. $[\,\cdot\,]_{\Lambda_\mb}$ denotes equivalence classes with respect to $Q_\mb^*(E_{\Lambda_\mb})$. \bigskip \begin{lem} \label{07075} Let $T$ be a substitution\ tiling with locally invertible substitution\ and $\Lambda_\mb$ and $Q_\mb$ as above. Then $E_{\cal R}/Q_\mb^*(E_{\Lambda_\mb})$ is generated by elements of the form $\left[\eta({\cal V}_\upsilon^n(A,F))\right]_{\Lambda_\mb}$ for $(A,F)\in{\cal B}^\upsilon_{2\neq}$. \end{lem} {\em Proof:} Since ${\mTxx}_{2,\neq}$ generates ${\cal M}_{\rm I\!I}\backslash{\mTxx}_{1}$ a generating set for $E_{\cal R}$ is provided by the set of elements of the form $\eta(\chi^{}_{c})$ where $c=uA$, $u\in{{\mTxx}^0}$, $u\vdash A\in{\mTxx}_{2,\neq}$. For such a $c$ let \begin{eqnarray*} I_1^{n}(c)&:=&\{(a,f,e_1,e_2)|(a,f)\in{\cal B}^\upsilon_1,e_i\in E^{n}(a)| \exists c':c, e_1 \hat{\rho}^{n-\upsilon}(f) e_2^{-1}\preceq c'\}\\ I_{2\neq}^{n}(c)&:=&\{(A,F,e_1,e_2)|(A,F)\in{\cal B}^\upsilon_{2\neq},e_i\in E^{n}(x_i(A))| \exists c':c, e_1 \hat{\rho}^{n-\upsilon}(F) e_2^{-1}\preceq c'\} \end{eqnarray*} Consider $\omega\in U_{L(c)}$ and let $n\geq \upsilon$. By Lemma~\ref{18061} there is a unique $(a,f,e_1)$ such that $\omega\in{\cal V}_\upsilon^n(a,f,e_1,e_1)$. It follows that $e_1^{-1}\vdash c$. Now suppose that $\exists e_2\in E^n(a)\exists c':e_1e_2^{-1}, c\preceq c'$. Then this $e_2$ is uniquely determined and $[\omega,c]\in {\cal V}_\upsilon^n(a,f,e_1,e_2)$. In particular ${\cal V}_\upsilon^n(a,f,e_1,e_2)$ are for $(a,f,e_1,e_2)\in I_1^{n}(c)$ pairwise disjoint. If the above assumption is not satisfied then, because of the form $c=uA$, there must be a $(A,F)$ with $a=L(A)$ and $f\preceq L(F)$ such that $\exists e_2\in E^n(x_2(A))\exists c':e_1\hat{\rho}^{n-\upsilon}(F)e_2^{-1}, c\preceq c'$. Again $e_2$ is uniquely determined and $[\omega,c]\in {\cal V}_\upsilon^n(A,F,e_1,e_2)$. This shows that ${\cal V}_\upsilon^n(A,F,e_1,e_2)$ are for $(A,F,e_1,e_2)\in I_{2\neq}^{n}(c)$ pairwise disjoint and hence \begin{equation} \label{30064} {\cal U}_c \subset \dot{\bigcup_{I^n_1(c)}}{\cal V}_\upsilon^n(a,f,e_1,e_2)\:\dot{\cup}\: \dot{\bigcup_{I_{2\neq}^n(c)}}{\cal V}_\upsilon^n(A,F,e_1,e_2). \end{equation} Since for all $(x,y)\in{\cal B}^\upsilon_1\cup{\cal B}^\upsilon_{2\neq}$ and $e_i\in E^{n}(x_i(x))$, $\mbox{\rm rad}(L(e_1\hat{\rho}^{n-\upsilon}(y)))$ diverges there is an $n$ such that for all $y$ and all $e_i$ either $c\preceq e_1\hat{\rho}^{n-\upsilon}(y)e_2^{-1}$ or $\not\exists c':c, e_1 \hat{\rho}^{n-\upsilon}(y) e_2^{-1}\preceq c'$. In other words, either ${\cal U}_c\cap{\cal V}^n_\upsilon(x,y,e_1,e_2)={\cal V}^n_\upsilon(x,y,e_1,e_2)$ or that intersection is empty. Thus the inclusion in (\ref{30064}) is an equality for large enough $n$. By (\ref{08071}), ${\cal V}_\upsilon^n(a,f,e_1,e_2)\in Q_\mb^{-1}({\cal ASG}({\cal R}_{\Lambda_\mb}))$ so that $[\eta({\cal V}_\upsilon^n(a,f,e_1,e_2))]_{\Lambda_\mb}=0$. Thus the first part of the union (\ref{30064}) will not contribute. As for the second, observe that \begin{equation}\label{08072} L({\cal V}_\upsilon^n(A,F,e_1,e_2))=\{\omega\cdot e_1^{-1}| \omega\in\rho^n(U_{L(A)})\cap\rho^{n-\upsilon}(U_{L(F)})\} \end{equation} and ${\cal V}_\upsilon^n(a,f,e,e)=\{\omega\cdot e^{-1}| \omega\in\rho^n(U_{a})\cap\rho^{n-\upsilon}(U_{f})\}$. As above, if $l$ is big enough one gets the alternative \begin{equation}\label{30063} L({\cal V}_\upsilon^n(A,F,e_1,e_2))\cap {\cal V}^{n+l}_1(a,f,e,e)=\left\{ \begin{array}{l} L({\cal V}_\upsilon^n(A,F,e_1,e_2)) \\ \emptyset \end{array}\right.. \end{equation} It follows that (\ref{08072}) (which does not depend on $e_2$) can be written as a disjoint union $$L({\cal V}_\upsilon^n(A,F,e_1,e_2))=\dot{\bigcup_{(a,f,e)\in J^l(A,F,e_1)}} {\cal V}^{n+l}_1(a,f,e,e)$$ for an appropriate $J^l(A,F,e_1)$, and moreover, $J^l(A,F,ce_1)$ with $R(c)=L(e_1)$ equals to $\{(a,f,ce)|(a,f,e)\in J^l(A,F,e_1)\}$. Thus $[L({\cal V}_\upsilon^n(A,F,e_1,e_2))]_{\Lambda_\mb}=[L({\cal V}_\upsilon^n(A,F))]_{\Lambda_\mb}$ and we end up with \begin{equation} [\eta({\cal U}_c)]_{\Lambda_\mb} = \sum_{(A,F,e_1,e_2)\in I_{2\neq}^n(c)}[\eta(V^n_\upsilon(A,F))]_{\Lambda_\mb}. \end{equation} Since $\eta(V^n_\upsilon(A,F))\in E_{\cal R}$ the lemma is proven.\hfill$\Box$ \begin{lem} Let $(a,f)\in{\cal B}^\upsilon_1$, $(A,F)\in{\cal B}^\upsilon_{2\neq}$, $e\in E^{l}(a)$, $n\geq \upsilon$, and $l\geq \upsilon$ large enough for the alternative (\ref{30063}) to hold. Then $L({\cal V}_\upsilon^n(A,F)) \cap {\cal V}_\upsilon^{n+l}(a,f,\hat{\rho}^n(e),\hat{\rho}^n(e))\neq\emptyset$ whenever $L(\hat{\rho}^\upsilon(e^{-1})F)\preceq\hat{\rho}^l(f)$ and $L(e^{-1}A)\preceq\hat{\rho}^{l-\upsilon}(f)$. \end{lem} {\em Proof:} $$ L({\cal V}_\upsilon^n(A,F)) \cap {\cal V}_\upsilon^{n+l}(a,f,\hat{\rho}^n(e),\hat{\rho}^n(e))\neq\emptyset $$ is for $n\geq \upsilon$ equivalent to $$ \{\omega\cdot \hat{\rho}^\upsilon(e) |\omega\in\rho^\upsilon(U_{L(A)})\cap U_{L(F)}\} \cap \rho^{l+\upsilon}(U_a)\cap \rho^l(U_f) \neq\emptyset. $$ That the latter implies $L(\hat{\rho}^\upsilon(e^{-1})F)\preceq\hat{\rho}^l(f)$ and $L(e^{-1}A)\preceq\hat{\rho}^{l-\upsilon}(f)$ is clear provided $l$ is large enough. Now let $(a,f)\in{\cal B}^\upsilon_1$, $(A,F)\in{\cal B}^\upsilon_{2\neq}$, and $\omega\in \rho^{l+\upsilon}(U_a)\cap \rho^l(U_f)$. Then $L(\hat{\rho}^\upsilon(e^{-1})F)\preceq \hat{\rho}^l(f)$ implies $L(F)\preceq \omega\cdot\hat{\rho}^\upsilon(e^{-1})\in \mbox{\rm im}\,\rho^\upsilon$, and $L(e^{-1}A)\preceq\hat{\rho}^{l-\upsilon}(f)$ implies $L(A)\preceq \rho^{-\upsilon}(\omega\cdot\hat{\rho}^\upsilon(e^{-1}))$. \hfill$\Box$\bigskip Define the $|{\cal B}^\upsilon_1|\times |{\cal B}^\upsilon_{2\neq}|$ matrices with coefficients $K^{(l,\mb)}_{(a,f)(A,F)}$ resp.\ $\KK^{(l,\mb)}_{(a,f)(A,F)}$ through \begin{equation} \label{07021} K^{(l,\mb)}_{(a,f)(A,F)}:= |\{e\in E^{l}(a)| L(\hat{\rho}^\upsilon(e^{-1})F)\preceq\hat{\rho}^l(f), L(e^{-1}A)\preceq\hat{\rho}^{l-\upsilon}(f) \}|. \end{equation} with $l\geq \upsilon$ and \begin{equation} \KK^{(l,\mb)}_{(a,f)(A,F)}:=K^{(l,\mb)}_{(a,f)(A,F)}-K^{(l,\mb)}_{(a,f)(A^{-1},F^{-1})}. \end{equation} \begin{thm} Let $T$ be a substitution\ tiling with locally invertible substitution\ and $\lambda_\mb$ and $\KK^{(l,\mb)}$ as above. If $l$ is large enough so that the alternative (\ref{30063}) holds then, with $k=|{\cal B}^\upsilon_1|$, \begin{equation} H({\cal R})\cong \{x\in \mbox{\rm \twlset R}^k|\exists n\geq 0:\lambda_\mb^n(x)\in\mbox{\rm \twlset Z}^k\}/ \{x\in \mbox{\rm \twlset R}^k|\exists n\geq 0:\lambda_\mb^n(x)\in\mbox{\rm im}\,\KK^{(l,\mb)}\}. \end{equation} \end{thm} {\em Proof:} By the last lemma $$ [\chi^{}_{L({\cal V}_\upsilon^n(A,F))}]_{\Lambda_\mb} = \sum_{(a,f)\in{\cal B}^\upsilon_1} [\chi^{}_{{\cal V}_\upsilon^{n+l}(a,f)}]_{\Lambda_\mb} K^{(l,\mb)}_{(a,f)(A,F)} $$ provided alternative (\ref{30063}) holds, and hence \begin{equation} \label{07071} [\eta({{\cal V}_\upsilon^n(A,F)})]_{\Lambda_\mb} = \sum_{(a,f)\in{\cal B}^\upsilon_1} [\chi^{}_{{\cal V}_\upsilon^{n+l}(a,f)}]_{\Lambda_\mb} \KK^{(l,\mb)}_{(a,f)(A,F)}. \end{equation} As we already saw, $[\chi^{}_{{\cal V}_\upsilon^{n}(a,f)}]_{\Lambda_\mb}=[Q_\mb^*(\chi^{}_\xi)]_{\Lambda_\mb}$ with $|\xi|=n$ and $r(\xi)=(a,f)$ so that these elements generate $H({\cal R}_{\Lambda_\mb})$. Moreover \begin{equation} \label{07022} [\chi^{}_{{\cal V}_\upsilon^{n}(a,f)}]_{\Lambda_\mb} = \sum_{(a',f')} [\chi^{}_{{\cal V}_\upsilon^{n+1}(a,f)}]_{\Lambda_\mb} {\lambda_\mb\,}_{(a',f')(a,f)} \end{equation} so that \begin{equation} \label{30062} {\cal L}^{(l+n,\upsilon)}=\lambda_\mb^n \KK^{(l,\mb)}. \end{equation} Recall (\ref{17061}) that $$H({\cal R}_{\Lambda_\mb}) = \{x\in \mbox{\rm \twlset R}^k|\exists n\geq 0:\lambda_\mb^n(x)\in\mbox{\rm \twlset Z}^k\}/ \{x\in \mbox{\rm \twlset R}^k|\exists n\geq 0:\lambda_\mb^n(x)=0\}.$$ The image of $E_{\cal R}/Q_\mb^*(E_{\Lambda_\mb})$ in $H({\cal R}_{\Lambda_\mb})$ is by Lemma~\ref{07075} and (\ref{07071},\ref{30062}) generated by the images of $j_n\circ\KK^{(l,\mb)}=\pi\circ\lambda_\mb^{-n}\circ\KK^{(l,\mb)}$. Now $\pi(x)\in\mbox{\rm im}\, j_n\circ\KK^{(l,\mb)}$ whenever $\lambda_\mb^n(x)\in\mbox{\rm im}\,\KK^{(l,\mb)}$ for some representative $x$ of $\pi(x)$. \hfill$\Box$\bigskip Suppose that $\Omega$ carries an ${\cal R}$-invariant (normalized) measure and that the substitution\ is primitive. Then this measure is unique (and the tiling minimal \cite{Ke2}) and the order structure of $H({\cal R})$ can be expressed as $[x]>0$ for $[x]\in H({\cal R})$ if and only if $\sum_i\nu_ix_i>0$ where $\nu$ is the left-Perron-Frobenius-ei\-gen\-vec\-tor\ of $\lambda_\mb$. In particular the latter inequality is independent of the chosen representative (the elements of $\ker[Q_\mb^*]$ are infinitesimal). \subsubsection{Simplifications} The determination of $\KK^{(l,\mb)}$ can be quite cumbersome. But simplifications occurr under certain circumstances. \medskip 1) In case that for all $(a,f)\in{\cal B}^\upsilon_1$, $a$ is uniquely determined by $f$ (we write it as $a(f)$) $L(\hat{\rho}^\upsilon(e^{-1})F)\preceq \hat{\rho}^l(f)$ implies $L(e^{-1}A(F))\preceq\hat{\rho}^{l-\upsilon}(f)$. Thus $K^{(l,\mb)}$ simplifies to \begin{equation} K^{(l,\mb)}_{fF} := K^{(l,\mb)}_{(a,f)(A,F)} = |\{e\in E^{l}(a(f))|L(\hat{\rho}^\upsilon(e^{-1})F)\preceq \hat{\rho}^l(f)\}|. \end{equation} This is for instance the case if $\upsilon=0$. 2) In some sense the other extreme is that the $f$ is determined by $a$. This is case if the substitution\ forces its border. It has been partly analyzed in \cite{Ke2}. In the present notation it is expressed as follows: \begin{df} A locally invertible substitution\ forces its border if there is a $\upsilon$ such that $\beta_\upsilon:{\cal B}^\upsilon_1\to{\mTxx}_{1}$ is a bijection. \end{df} Not only the graph simplifies enormously, in that it coincides with the improper graph, but also we may take $l=\upsilon$. Moreover $L(e^{-1}A)\preceq\hat{\rho}^{l-\upsilon}(f(a))$ implies $L(\hat{\rho}^\upsilon(e^{-1})F(A))\preceq\hat{\rho}^l(f(a))$ where we wrote $f(a)$ for the $f$ determined by $a$. Hence for substitution s which force their border \begin{equation} K^{(\upsilon,\upsilon)}_{aA}:= K^{(\upsilon,\upsilon)}_{(a,f)(A,F)} = |\{e\in E^{\upsilon}(a)|L(e^{-1}A)\preceq f(a)\}|. \end{equation} Note that a substitution\ can force its border only for $\upsilon\geq 1$, because otherwise any pointed tiling would be determined by its pointed tile and hence periodic which contradicts local invertibility. \medskip As for the determination of $H({\cal R})$ we have:\medskip 3) If $\lambda_\mb$ is invertible over $\mbox{\rm \twlset Z}$ then $H({\cal R})=\mbox{\rm \twlset Z}^k/\langle\bigcup_n \lambda_\mb^{-n}\mbox{\rm im}\,\KK^{(l,\mb)}\rangle$. \subsection{Examples} The above machinerie has been designed to tackle higher dimensional tilings. We therefore will present the computation of the coinvariants associated to the Penrose tilings. But it is also applicable to one dimensional substitution s in the sense of \cite{Q}. To allow the reader a comparison with technics used elsewhere to obtain the integer group of coinvariants, c.f.\ \cite{For,Hos}, we present the Thue Morse substitution\ as an example. \subsubsection{The Thue Morse substitution} The Thue Morse substitution\ $\varrho$ is defined on the two letter alphabet $\{a,b\}$ by \begin{eqnarray} a &\mapsto& ab\\ b &\mapsto& ba \end{eqnarray} and extended to words as $\varrho(a_1\dots a_k)= \varrho(a_1)\dots \varrho(a_k)$. It may be viewed (like any other one dimensional substitution of the kind in \cite{Q}) as a substitution\ in the algebraic sense of the tiling which is a fixed point under $\varrho$: Consider the sequence over $\mbox{\rm \twlset Z}^{\geq 0}$ with values in $\{a,b\}$ given by $\varrho^\infty(a)$ and complete it to a sequence over $\mbox{\rm \twlset Z}$ by reflection (i.e.\ by $\varrho^{2\infty}(a)$ to the left). This represents a one dimensional pointed tiling the pointed tile (letter) being the one on $0$, i.e.\ the first one of (the right) $\varrho^\infty(a)$. The (finite) words appearing in $\varrho^\infty(a)$ with two letters chosen are doubly pointed pattern class es.\footnote{ If we do not have a geometric interpretation of the tilings as sequences of decorated intervals we have to restrict to pattern class es which are connected.} We indicate the first by a grave and the second by an acute, and if both coincide by a check. (An example of a multiplication is $(\lp{a}\fp{b})(b\fp{a}b\lp{b})=(ba\fp{b}\lp{b}))$. A substitution\ as a homomorphism of the almost-groupoid\ of doubly pointed pattern class es is then given by $\hat{\rho}(w_1\fp{a}_iw_2\lp{a}_jw_3)=\varrho(w_1)\fp{a}_{i_1}a_{i_2} \varrho(w_2)\lp{a}_{j_1}a_{j_2}\varrho(w_3)$ where we have used the writing $\varrho(a_i)=a_{i_1}a_{i_2}$. The substitution\ is locally invertible but does not force its border and there is no advantage in using large $\upsilon$. We therefore take $\upsilon=0$. The connectivity matrix\ of $\Sigma_0$ is $${\lambda_0\,}_{f_1f_2} =\mbox{number of $e\in E^1(a(f_1))$ with $f_2\preceq L(e\check{\rho}(f_1))$}. $$ We have $${\cal B}^0_1=\{a\bp{a}b,b\bp{a}a,b\bp{a}b,b\bp{b}a,a\bp{b}b,a\bp{b}a\}$$ $$\hat{\rho}({\cal B}^0_1)=\{ab\bp{a}bba,ba\bp{a}bab,ba\bp{a}bba, ba\bp{b}aab,ab\bp{b}aba,ab\bp{b}aab\}$$ $$E^1(a)=\{\bp{a}b,\lp{a}\fp{b}\}\quad E^1(b)=\{\bp{b}a,\lp{b}\fp{a}\}$$ Taking the elements of ${\cal B}^0_1$ and $\hat{\rho}({\cal B}^0_1)$ in the above order one obtains $$\lambda_0 = \left(\begin{array}{cccccc} 0 & 0 &1 &0 &1 &0 \\ 1 & 0 &0 &0 &0 &1 \\ 1 & 0 &0 &0 &1 &0 \\ 0 & 1 &0 &0 &0 &1 \\ 0 & 0 &1 &1 &0 &0 \\ 0 & 1 &0 &1 &0 &0 \end{array}\right). $$ The Perron-Frobenius-eigen\-value\ is $2$ and the normalized left-Perron-Frobenius-ei\-gen\-vec\-tor\ is $\nu=\frac{1}{6}(1,1,1,1,1,1)$. Hence \begin{equation} \mu(H({\cal R})) = \frac{1}{3}\mbox{\rm \twlset Z}[\frac{1}{2}]. \end{equation} To obtain the full group we may split up ${\cal B}^0_{2\neq}={\cal B}^0_{2<}\cup ({\cal B}^0_{2<})^{-1}$ with $${\cal B}^0_{2<}=\{a\fp{a}\lp{b}a,a\fp{a}\lp{b}b,b\fp{a}\lp{b}a,b\fp{a}\lp{b}b, b\fp{a}\lp{a}b,b\fp{b}\lp{a}b,b\fp{b}\lp{a}a,a\fp{b}\lp{a}b,a\fp{b}\lp{a}a, a\fp{b}\lp{b}a\}.$$ Thus (\ref{30063}) holds for $l=1$, i.e.\ we have to determine $$K^{(1,0)}_{fF} =\mbox{number of $e\in E^1(a(f))$ with $L(F)\preceq L(e\hat{\rho}(f))$}. $$ Let $ {\cal L}^{(1,0)}_{<}$ be the restriction of ${\cal L}^{(1,0)}$ to indices $fF$ with $F\in{\cal B}^0_{2<}$. Then, again with respect to the above order, $$ {\cal L}^{(1,0)}_{<}= \left(\begin{array}{cccccccccc} 0 & 0 &0 &0 &0 &0 &0 &-1 &0 &1\\ 0 & 0 &0 &0 &-1 &0 &0 &1 &0 &0\\ 0 & 0 &0 &0 &-1 &0 &0 &0 &0 &1\\ 0 & 0 &-1 &0 &1 &0 &0 &0 &0 &0\\ 0 & 0 &1 &0 &0 &0 &0 &0 &0 &-1\\ 0 & 0 &0 &0 &1 &0 &0 &0 &0 &-1 \end{array}\right). $$ And, since ${\cal L}^{(1,0)}_{fF^{-1}} = - {\cal L}^{(1,0)}_{fF}$, $\mbox{\rm im}\,{\cal L}^{(1,0)}=\mbox{\rm im}\,{\cal L}^{(1,0)}_{<}\cong\mbox{\rm \twlset Z}^3$. $\lambda_0{\cal L}^{(1,0)}_{<}$ is up to a permutation of the columns ${\cal L}^{(1,0)}_{<}$, i.e.\ $\lambda_0$ preserves $\mbox{\rm im}\,{\cal L}^{(1,0)}$. In fact, $\lambda_0$ is diagonalizable, it has eigenvalues $2,1,0,-1$, $-1$ occurring with multiplicity $3$, and $\mbox{\rm im}\,{\cal L}^{(1,0)}$ is spanned by the (right) eigenvector to eigenvalue $1$ together with a two dimensional subspace of the eigenspace to eigenvalue $-1$. A system of (right) eigenvectors is given by $$ \left(\begin{array}{c} 1\\1\\1\\1\\1\\1 \end{array}\right)\quad \left(\begin{array}{c} -1\\0\\-1\\1\\0\\1 \end{array}\right)\quad \left(\begin{array}{c} -1\\-1\\-1\\1\\1\\1 \end{array}\right)\quad \left(\begin{array}{c} 0\\-1\\-1\\0\\1\\1 \end{array}\right)\quad \left(\begin{array}{c} -1\\1\\0\\-1\\1\\0 \end{array}\right)\quad \left(\begin{array}{c} 1\\-1\\-1\\1\\0\\0 \end{array}\right). $$ It follows that \begin{equation} H({\cal R})\cong\mbox{\rm \twlset Z}[\frac{1}{2}]\oplus\mbox{\rm \twlset Z}. \end{equation} The elements of the second summand are infinitesimal since the pairing between the left Perron-Frobenius-ei\-gen\-vec\-tor\ $\nu$ and all vectors from the (right) eigenspace to eigenvalue $-1$ is zero. Thus the positive cone is $\mbox{\rm \twlset Z}[\frac{1}{2}]^+$. \subsubsection{Penrose tilings} \label{03081} There are several well known variants of tilings which are called Penrose tilings \cite{Pen,GrSh} and which are a priori to be distinguished as they lead to non-isomorphic groupoids. But they may be transformed into each other by purely local manipulations which implies that one can find maps satisfying the conditions of Lemma~\ref{27061} and leading to isomorphisms between reductions of the corresponding groupoids. Since any such tiling is minimal all reductions lead to the same ordered integer group of coinvariants differing possibly in the order unit. The version which is most suitable for our purposes is the one which has triangles as tile s, cf.\ Figure~7. The triangles are decorated (with a little circle) to break the mirror symmetry. There are $40$ pattern class es of them. The orientational symmetry of a tiling (or its class) is the largest subgroup of $O(d)$, acting on a pointed tiling being identified with a representative in $\mbox{\rm \twlset R}^d$ in the obvious way, which leaves the hull invariant. The orientational symmetry of a Penrose tiling by triangles possesses $20$ elements. It is generated by a rotation around $\frac{\pi}{5}$ together with a mirror reflection at a boundary line of a triangle \cite{Ke2}. We denote by ${\cal M}_{\rm I\!I}$ resp.\ ${\cal R}$ the almost-groupoid\ resp.\ groupoid associated to these tilings. The well known deflation of these tilings $\rho_t$ with $t=\frac{1+\sqrt{5}}{2}$ is displayed below. Since it is covariant with respect to the orientational symmetry it suffices to give it for one orientation only. \epsffile[0 0 430 170]{dFig1.ps} A replacement $\rho_t(a)$ taking exactly the same space as the tile $a$ the relative position measured with respect to the centers of gravity is $0$. Choosing pointed tiles for $\rho_t(a)$ we obtain a substitution\ $\hat{\rho}$. Those tilings which have an exact five-fold symmetry are invariant under $\hat{\rho}^4$. The substitution\ is primitive, locally invertible, and forces its border with $\upsilon=4$ \cite{Ke2}. Hence $\Lambda_4=\Sigma$ and $\lambda_4=\sigma$, the substitution\ matrix of $\hat{\rho}$. Since $\sigma$ is invertible over the integers $H({\cal R}_\Sigma)\cong\mbox{\rm \twlset Z}^{40}$. To simplify the computation of $E_{\cal R}/Q_4^*(E_\Sigma)$ we make use of the symmetry properties of the tiling and in particular of the fact that the boundaries of substitute s $\hat{\rho}^4(a)$ are local mirror axes so that the pattern classes of those $A\in{\mTxx}_{2,\neq}$ which cross the boundaries of $4$-fold substitute s are always mirror symmetric \cite{Ke2}, c.f.\ below where the boundaries are indicated through fatter lines. \epsffile[0 0 430 190]{dFig7.ps} Let $\alpha$ denote the direction of a boundary line of a tile resp.\ a substitute. There are $10$ different ones and we order them anti-clockwise identifying them with $\{0,\cdots,9\}$. Saying that a substitute\ has boundary $\alpha$ if it has a boundary with that direction we define the $40\times 40$ matrices with entries \begin{equation} N^\alpha_{ab} := \mbox{number of $a$'s at boundary $\alpha$ of $\hat{\rho}^4(b)$} . \end{equation} In particular $N^\alpha_{ab}=0$ in case $a$ or $\rho(b)$ do not have boundary $\alpha$. Let $\alpha(a)$ be the mirror image of $a$ with respect to the mirror axis $\alpha$ and define \begin{equation} {\cal D}^\alpha_{ab} = N^\alpha_{ab}-N^\alpha_{a \alpha(b)}. \end{equation} Then \begin{equation} {\cal L}^{(4,4)}_{a A} = \left\{ \begin{array}{cl} {\cal D}^\alpha_{ab} & \mbox{if } |A|=b \alpha(b)\mbox{ and } x(A)=b \\ -{\cal D}^\alpha_{ab} & \mbox{if } |A|=b \alpha(b)\mbox{ and }x(A)=\alpha(b)\\ 0 & \mbox{else} \end{array} \right. \end{equation} $|A|=b\alpha(b)$ indicating that the pattern class of $A$ is composed of $b$ and $\alpha(b)$ in such a way that the common boundary is the symmetry axis. Hence $\mbox{\rm im}\, {\cal L}^{(4,4)}=\erz{\mbox{\rm im}\,{\cal D}^\alpha,\alpha=0,\dots,9}$. ${\cal D}^\alpha$ is related to ${\cal D}^{0}$ by symmetry, i.e.\ ${\cal D}^\alpha=R^{-\alpha}{\cal D}^{0} R^\alpha$, $R$ being the matrix which acts as a rotation around $\frac{\pi}{5}$. To be very explicit let us use a basis $\{e_{10 k+\alpha}\}_{0\leq k \leq 3, 0\leq\alpha\leq 9}$ of $H({\cal R}_\Sigma)$ with $e_{10 k+\alpha}=[\chi^{}_{U_{a_{10 k+\alpha}}}]_\Sigma$ where $a_{10 k+\alpha}$ corresponds to the pattern class\ of the triangle in Figure~7.k rotated around an angle of $\frac{\alpha\pi}{5}$. \epsffile[0 0 430 100]{dFig2.ps} In terms of the rotation matrix $\omega$, which has entries $\omega_{\alpha\beta}=\delta_{\alpha-\beta,1\;mod\;10}$, $R$ and the substitution\ matrix $\sigma$ are given by $$ R = \left( \begin{array}{cccc} \omega & 0 & 0 & 0 \\ 0 & \omega & 0 & 0 \\ 0 & 0 & \omega & 0 \\ 0 & 0 & 0 & \omega \end{array} \right), \quad \sigma = \left( \begin{array}{cccc} \omega^4 & \omega^0 & 0 & \omega^6 \\ \omega^0 & \omega^6 & \omega^4 & 0 \\ \omega^3 & 0 & \omega^7 & 0 \\ 0 & \omega^7 & 0 & \omega^3 \end{array} \right). $$ The matrix $N^{0}$ may be read of from Figure~8 (and completed by symmetry); it is given below (\ref{19011}). Moreover ${\cal D}^0=N^0-N^0S$ where $S$ implements the reflection at $\alpha=0$, explicitly, with $s_{\alpha\beta}=\delta_{\alpha+\beta,5\;mod\;10}$ (counting rows and columns form $0$ to $9$) $$ S = \left( \begin{array}{cccc} 0 & s & 0 & 0 \\ s & 0 & 0 & 0 \\ 0 & 0 & 0 & s \\ 0 & 0 & s & 0 \end{array} \right). $$ It turns out that $\mbox{\rm im}\,{\cal L}^{(4,4)}$ is generated by the orbit under $R$ of (the transpose of) the four vectors $$\begin{array}{c} v_1=(0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0, \!0,\!0,\!0,\!0,\!0,\! 0,\!0,\!1,\!0,\!0,\!0,\!0,\!-1,\!0,\!0,\!0,\!0, \!0,\!0,\!0,\!0)\\ v_2=(0,\!0,\!0,\!0,\!0,\!0,\!1,\!0,\!0,\!0,\!0,\!-1,\!0,\!0,\!0, \!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\! 0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0)\\ w_1=(0,\!0,\!0,\!0,\!0,\!0,\!0,\!1,\!0,\!0, -1,\!0,\!0,\!0,\!0, \!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\! 0,\!0,\!0,\!0,\!1,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0, -1)\\ w_2=(0,\!0,\!0,\!0,\!0,\!1,\!0,\!0,\!0,\!0,\!0,\!0, -1,\!0, \!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\! 0,\!0,\!0,\!0,\!0, -1,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!0,\!1,\!0) \end{array} $$ but $8$ of the $40$ vectors thus obtained are linearly dependent of the others. Moreover $\mbox{\rm im}\,{\cal L}^{(4,4)}$ is invariant under $\sigma$. Dividing it out yields no torsion so that we obtain \begin{equation}\label{31071} H({\cal R})\cong\mbox{\rm \twlset Z}^8. \end{equation} To obtain the order structure we look for the smallest subgroup $I$ of $\mbox{\rm \twlset Z}^{40}$ which is invariant under $\sigma$ and spans a real vector space containing the right Perron-Frobenius-ei\-gen\-vec\-tor\ of $\sigma$. Then the positive cone of $\mbox{\rm \twlset Z}^{40}$ is given by those $x\in I$ which pair with the left-Perron-Frobenius-ei\-gen\-vec\-tor\ $\nu$ of $\sigma$ to $\nu x >0$. All other elements must pair with $\nu$ to zero and are thus infinitesimal. Clearly $I$ is spanned by (the transpose of) $$ x_1 = (\overbrace{1,\dots,1}^{20},\overbrace{0,\dots,0}^{20}), \quad x_2 = (\overbrace{0,\dots,0}^{20},\overbrace{1,\dots,1}^{20}) $$ and $\nu$ is given by $\nu = \frac{2-t}{20}(tx_1+x_2)^T$ (in its normalized form). Thus $I\cap \mbox{\rm im}\,{\cal L}^{(4,4)}=\emptyset$ and \begin{equation}\label{03083} H({\cal R})\cong I \oplus \mbox{\rm \twlset Z}^6 \end{equation} the elements of the second summand being infinitesimal. The range of the unique state on $H({\cal R})$, which of course coincides with that on $H({\cal R}_\Sigma)$, is \begin{equation} \label{03082} \mu(H({\cal R}))=\frac{1}{20}(\mbox{\rm \twlset Z} + t\mbox{\rm \twlset Z}). \end{equation} This concludes the computation of the integer group of coinvariants and the range of its state for the Penrose tilings. But it may be instructive to look at result (\ref{31071}) from a different point of view. Let $H_1$ be the sublattice of $\mbox{\rm \twlset Z}^{40}$ which is generated by $\{R^\alpha v_i\}_{i=1,2;\,\alpha=0,1\cdots,9}$. $H({\cal R}_\Sigma)/H_1=\mbox{\rm \twlset Z}^{20}$ has a basis with natural geometrical interpretation. It is formed by the classes of characteristic functions on the pattern class es of rhombi which are always formed by either two smaller triangles or two larger triangles. This indicates that one should look at the Penrose tilings by rhombi. Such a tiling, which we denote by $T_{Rh}$, has only $20$ pattern class es of tile s. That the groupoid associated to the Penrose tilings by rhombi is isomorphic to a reduction of ${\cal R}$ may be seen as follows: Let ${\cal C}\subset{\mTxx}_{1}$ consist of the classes of triangles that are obtained from those of Figure~7.1 and 7.3 by a rotation. Let ${Rh}\subset{\cal M}_{\rm I\!I}$ be the almost-groupoid\ given by elements of the form $aca'$ where $a,a'\in{\cal C}$ and $c\in{\cal M}_{\rm I\!I}$ is such that all its triangles pair to rhombi. Clearly ${Rh}$ is approximating generating and regular. Now deleting the diagonal which coincides with the base of the two touching triangles yields an isomorphism of almost-groupoid s ${Rh}\to\mtxx{T_{Rh}}$ which satifies the growth condition with $t=1$. Hence the groupoid associated to the rhombus version is isomorphic to ${\cal R}_{Rh}$. The largest subgroup of $O(2)$ leaving $\Omega_{Rh}$ invariant consists only of the rotations which are multiples of $\frac{\pi}{5}$ since the mirror image of a $\omega\in\Omega_{Rh}$ lies in $\Omega\backslash\Omega_{Rh}$. However, to our knowledge there is no substitution\ for Penrose tilings by rhombi which is covariant even under this reduced orientational symmetry. But there are non covariant ones, namely $10$ of them, the deflation corresponding to $\hat{\rho}_0$ being given in Figure~9 and $\hat{\rho}_\alpha$ being obtained from $\hat{\rho}_0$ just by rotation of the whole figure around $\frac{\alpha\pi}{5}$, $\alpha=0,\cdots,9$. (The relative positions are indicated by a cross.) That all these substitution s are primitive locally invertible and force their border\ carries over from $\hat{\rho}$. It turns out that \begin{equation} \ker\sigma_\alpha = \erz{\{[R^\alpha w_i]_{H_1}\}_{i=1,2}} \end{equation} where $\sigma_\alpha$ is the substitution\ matrix of $\hat{\rho}_\alpha$ and $[\,\cdot\,]_{H_1}$ denotes the classes in $H({\cal R}_\Sigma)/H_1$. Since the restriction of $\sigma_\alpha$ to its image is an automorphism and $\ker\sigma_0=\ker\sigma^2_0$ we have \begin{equation} H({\cal R}_{\Sigma_\alpha}) \cong H({\cal R}_\Sigma)/\erz{\{H_1,R^\alpha w_i,i=1,2\}} \end{equation} $\Sigma_\alpha$ denoting the graph having $\sigma_\alpha$ as connectivity matrix. Any of the substitution s $\hat{\rho}_\alpha$ leads to a homeomorphism $Q_\alpha$ between $\Omega_{Rh}$ and the path space ${\cal P}_{\Sigma_\alpha}$ and to a surjective homomorphism $[Q^*_\alpha]:H({\cal R}_{\Sigma_\alpha})\rightarrow H({\cal R})$. Let $\pi_\alpha:H({\cal R}_\Sigma)\to H({\cal R}_{\Sigma_\alpha})$ be the natural projection. Then $[Q_4^*]=[Q_\alpha^*]\circ \pi_\alpha$ for all $\alpha$ and therefore \begin{equation} \ker [Q_4^*] \supset\erz{\{H_1,R^\alpha w_i,i=1,2;\,\alpha=1,\cdots,9\}}\cong\mbox{\rm \twlset Z}^{32}. \end{equation} This shows independently that $H({\cal R})\subset\mbox{\rm \twlset Z}^{8}$ but not the opposite inclusion. A computation of $\mbox{im}\,{\cal L}^{(4,4)}$ for e.g.\ $\rho_0$ would have been more complicated due to the lack of symmetry. Connes associates to the Penrose tilings yet another graph, the folded $A_4$ Coxeter graph, making use of a coding of a tiling by $0,1$ sequences obeying the condition that no consecutive $1$'s can appear \cite{Cone}. But this coding, which was found by Robinson \cite{GrSh}, does not distinguish between a tiling and its image under an element of the orientational symmetry. In fact, the coding yields a homeomorphism between the hull modulo the orientational symmetry and the path space of the folded $A_4$ graph, with the effect that the groupoid arising is that given by cofinality. In other words, the tilings are considered as equivalence classes under translations {\em and} rotations and reflections. One then obtains an $AF$-algebra as groupoid-$C^*$-algebra\ whose $K_1$-group is trivial and whose $K_0$-group coincides with the integer group of coinvariants of the groupoid and may be identified with the group $I$ in (\ref{03083}) as ordered group. Although the range of the tracial state on it coincides with (\ref{03082}) up to the order unit (the factor $\frac{1}{20}$ does not appear) it is a priori not clear that it predicts the right gap labels since the $AF$-algebra does not contain the discrete Laplacian. \subsection*{Concluding Questions} \addcontentsline{toc}{section}{\bf Concluding Questions} We have computed the integer group of coinvariants and its order but even in two dimensions and under the assumption that $K_0({\cal A}_T)=H({\cal R})\oplus \mbox{\rm \twlset Z}$ it is not clear what the order structure on the $K_0$-group is.\medskip We did not mention groupoid cohomology but at least for tilings which reduce to decorations the integer group of coinvariants is a cohomology group of the groupoid ${\cal R}$ with coefficients in $\mbox{\rm \twlset Z}$. The result of \cite{FoHu} on the connection between $K$-theory and group cohomology is easily seen to generalize to this situation since the stability of $K$-theory under taking stably isomorphic algebras is mirrored by the stability of (continuous) groupoid cohomology under taking (continuously) similar groupoids. It is therefore tempting to believe that $K$-theory of algebras associated to tilings is always related to the groupoid cohomology with coefficients in $\mbox{\rm \twlset Z}$ of the associated groupoid in a way like in \cite{FoHu}. \newpage \begin{equation} \label{19011} N^{0} = \left(\begin{array}{c} 3\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;1\;0\;2\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;1\;0\;0\;1\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 0\;1\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\;0\;2\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;1\;0 \;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;1\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 1\;0\;0\;0\;1\;0\;1\;0\;0\;0\;1\;0\;0\;0\;0\;0\;1\;0\;0 \;0\;1\;0\;0\;0\;0 \;0\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;3\;0\;0\;0\;0\;0\;2\;0\;0\;0\;1 \;0\;0\;0\;1\;0\;0\;1\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\\ 0\;0\;0\;0\;1\;0\;1\;0\;0\;0\;2\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\; 0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;1\;0\;0\;0\;1\;0\;0\;0\;1\;0\;1\;0\;0\;0\;1\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;1\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 1\;0\;0\;0\;2\;0\;1\;0\;0\;0\;3\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;1\;0\;0\;0\;0\;0 \;0\;0\;0\;0\; 0\;0\;0\;1\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;1\;0\;0\;0\;1\;0\;1\;0\;0\;0\;1 \;0\;0\;0\;1\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 2\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0 \;1\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;1\;0\;0\;1\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;1\;0\;0\;0\;1\;0\;0\;0\;2\;0\;0\;0\;0\;0\;3\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;1\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;1\;0\\ 1\;0\;0\;0\;1\;0\;0\;0\;0\;0\;1\;0\;0\;0\;1\;0\;1\;0\;0\;0 \;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\; 1\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 0\;0\;0\;0\;0\;2\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\;0\;0\;0\;1 \;0\;0\;1\;0\;0\;0 \;0\;0\;0\;0\; 0\;0\;0\;0\;0\;1\;0\;0\;0\;0\\ 1\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0 \;1\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 0\;1\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\;0\;2\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;1 \;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;1\;0\\ 1\;0\;0\;0\;1\;0\;0\;0\;0\;0\;1\;0\;0\;0\;1\;0 \;1\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 0\;0\;0\;0\;0\;1\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\;0\;0\;0\;1\;0 \;0\;0\;0\;0\;0\;0 \;0\;0\;0\; 0\;0\;0\;0\;0\;1\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;1\;0\;1\;0\;0\;0\;2\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0 \;0\;0 \;0\;0\;0\;0\;0\;0\;1\;0\;0\; 0\;0\;0\;0\\ 0\;0\;0\;0\;0\;1\;0\;0\;0\;1\;0\;1\;0\;0\;0\;1\;0\;0\;0\;1 \;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\; 0\;0\;0\;0\;0\;1\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;1\;0\;1\;0\;0\;0\;1\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0 \;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;1\;0\;0\;0\;1\;0\;0\;0\;1\;0\;1\;0\;0\;0\;1 \;0\;0\;0\;0\;0\;0\;0\;0\; 0\;1\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 2\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0\;1\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0 \;1\;0\;0\;1\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;1\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\;0\;1 \;0\;0\;0\;0\;0\;0\;0\;0\; 0\;1\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0\\ 1\;0\;0\;0\;1\;0\;1\;0\;0\;0\;1\;0\;0\;0\;0\;0\;1\;0\;0\;0 \;1\;0\;0\;0\;0\;0 \;0\;0\;0\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;2\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0 \;0\;0\;0\;1\;0\;0\;1\;0\; 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;1\;0\;0\;0\;0\\ 0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\;0 \;0\;0\;0\;0\;0\;0\;0\;0\;0\;0\; 0\;0\;0\;0 \end{array}\right) \end{equation} \newpage \epsffile[0 0 465 615]{dFig3.ps} Figure 8: $4$-fold substitutes of the triangles of Figure~7.0 and 7.1 in all directions. \newpage \epsffile[0 0 465 645]{dFig4.ps} \newpage \addcontentsline{toc}{section}{\bf References}
\section{Statement of the problem} The Dirac-ADM canonical approach \cite{Dir58,ADM} to GR was the essential improvement on the way to quantization of the Einstein-Hilbert theory of gravity developed by Wheeler, DeWitt and others \cite{Wheel,DeWitt,Mis,Ryan}. This conventional scheme of the canonical quantization is based on the (3+1) Dirac-ADM foliation \cite{Dir58,ADM,kuchar,Zel} of the four dimensional manifold $(x^\mu)$ along the some time-like vector (associated with the rest frame of an observer): $$ ds^2=N^2dt^2-{}^{(3)}g_{ik}\breve d x^i \breve d x^k \;\;;\;\; (\breve d x^i=dx^i + N^idt)\, $$ (that means the restriction of the group of general coordinate transformations by the kinemetric ones \cite{Zel}: $t\to t'(t)\;\;\;;\;\;\;x_i\to {x'}_i(t,x_1,x_2,x_3)$) and on the ADM action ($W_{ADM}$) which differs from the initial Einstein-Hilbert action \cite{Hil15} \begin{equation} \label{action} W_{GR}=\int d^4x\sqrt{-g}\left [-\frac{{}^{(4)}R(g)}{2\kappa^2} + {\cal L}_{matter} \right ]\;\;;\;\;(\kappa^2=8\pi G), \end{equation} by the surface terms: $$ W_{GR}=W_{ADM}+W_S+W_T, $$ where \begin{equation} \label{t} W_T=-\int dtd^3x {\breve \partial}_0 \left [\frac{{\breve \partial}_0 \sqrt{{}^{(3)}g}}{N\kappa^2} \right ]\;\;;\;\;\left [{\breve \partial}_0f= \dot f -\partial_k (N^kf)\right]. \end{equation} \begin{equation} \label{s} W_S=-\int dtd^3x \partial_k[\sqrt{{}^{(3)}g}({}^{(3)}g^{ik} \partial_l N)]\frac{1}{\kappa^2}\;. \end{equation} For the derivation of local classical equations these surface terms are not essential, however, they play an important role in the determination of the global quantities of such a total energy \cite{Regge}. It is obvious that the wave function also has global nature and depends on these surface terms. {\bf The statement of the problem consists in the canonical quantization of the Einstein-Hilbert action~(\ref{action}) by taking into account the surface terms (\ref{t}),~(\ref{s})}. We continue the attempts to solve this problem in the papers \cite{Perv,Khved}. \section{A new version of the Hamiltonian Formulation} {\bf To solve this problem we should I) consider the space-scale variable $a=[{}^{(3)}g]^{\frac{1}{6}}$ in the time surface term~(\ref{t})} as one of dynamical variables : \begin{equation} \label{metric} ds^2=a^2[N_c^2dt^2-\omega_{i\underline k}\omega_{j\underline k}{\breve d}x^i\breve d x^j], \;\;\;;\;\;\;aN_c=N\;\;\;;\;\;\; {\rm det}\omega=1, \end{equation} (we use the triad form $\omega_{i\underline j}$ for the rest dynamic variables~\cite{Smir}) {\bf II) apply the Ostrogradsky method} ~\cite{Ostr,gkp} to the theory with second order derivative of the scale factor with respect to the time coordinate. As the result, the Hilbert action (\ref{action}) in terms of the canonical conjugate variables $ a,\pi_a, \Phi,\pi_\Phi$ (where $\Phi$ denotes the set of matter fields including graviton $\omega$ and photon $A$) has the form $$ W_{GR}=\int^T_0 dt \int_V d^3 x \left [-\pi_{(a)}\stackrel{\odot}{a}+ \frac{1}{2}{\breve\partial}_0(\pi_{(a)}a)+\sum_{\Phi =\omega ,A} \pi_{(\Phi)} \stackrel{\odot}{\Phi} -N_c{\cal H}_{EC}\right] + W_S, $$ where $$ \stackrel{\odot}{a}=\breve\partial{}_0a+\frac{2}{3}a\partial_kN^k,\;\;\;\;\; \stackrel{\odot}{A}_k=\partial_0A_k-\partial_kA_0-N^lF_{lk} $$ $$ \stackrel{\odot}{\omega}_{lk}=(\dot\omega {}_{l\underline s}\omega {}_{k\underline s} +\omega {}_{l\underline s}\dot\omega {}_{k\underline s} -\nabla_lN_k -\nabla_kN_l+ \frac{2}{3}\omega_{l\underline s}\omega_{k\underline s}\partial_jN^j), $$ is the kinemetric invariant time derivative (we use here covariant derivative in the metric $\omega_{i\underline k}\omega_{j\underline k}$, including the Laplace operator $\Delta f=\nabla_k\partial^kf$), and ${\cal H}_{EC}$ is the Einstein energy density \begin{equation} \label{eenergy {\cal H}{}_{EC}=-\frac{\kappa^2}{12}\pi^2_{(a)}+ \pi^2_{(\omega)}\left( \frac{2\kappa^2}{a^2}\right)+\left(\frac{a^2}{2\kappa^2}\right)\bar R+ {\cal H}_{(A)};\;\;\;\; \end{equation} with the photon energy $$ {\cal H}{}_{(A)}= \frac{1}{2}\pi_{(A)}^k{\pi}_{k (A)}+\frac{1}{4}F_{ij}F^{ij} \;\; $$ and three dimensional curvature $$ \bar R=a^2\;\;{}^{(3)}R(a^2\omega^2)={}^{(3)}R(\omega^2) + 8a^{-\frac{1}{2}}\Delta a^{\frac{1}{2}}, $$ {\bf III) perform the canonical transformation} $(\pi_{(a)},a) \ =>\;(\Pi,\eta)$ which removes the time surface term $\frac{1}{2}\partial_0(\pi_{(a)}a)$ \begin{equation} \label{ct} \pi_{(a)}\stackrel{\odot}{a}-\frac{1}{2}\breve\partial{}_0(\pi_{(a)}a)= \Pi (\dot\eta-N^k\partial_k\eta)\;. \end{equation} One can represent this transformation as \begin{equation} \label{cantr} \pi_{(a)}=2\sqrt{\frac{3\Pi}{\kappa^2\Gamma}} C(\eta);\;\;\;\; a=\sqrt{\frac{\kappa^2\Gamma \Pi}{3}} S(\eta), \end{equation} where $ C(\eta), S(\eta) $ and $\Gamma $ are some particular solution of the following equations \begin{equation} \label{cs} C(\eta)\frac{d}{d\eta} S(\eta)- S(\eta)\frac{d}{d\eta} C(\eta)=1;\;\;\;\;\; \partial_0({\rm ln\Gamma})-N^k\partial_k({\rm ln}\Gamma)+\frac{1}{3}\partial_kN^k=0\;. \end{equation} Finally, the Hilbert action reads in terms of the new variables as \begin{equation} \label{ham} W_{GR}=\int dtd^3x\left [\sum_{\Phi =\omega ,A,{\rm ln}\Gamma } \pi_{(\Phi )} {\dot \Phi }-\Pi {\dot \eta } -N_c{\cal H}{}_{EC}-N^k{\cal P}_k \right]+{\bar W} {}_S\;, \end{equation} with space surface term $\bar W{}_S=W_S-2\int^T_0dtd^3x\partial_l(N^k\pi_{(h)}{}^l_k)$ and constraints \begin{equation} \label{c1} {\cal H}_{EC}= -\frac{\Pi}{\Gamma}(C^2-S^2\frac{6}{\Gamma^2}\bar R) +\pi^2_{(\omega)}\frac{6}{S^2\Pi\Gamma}+{\cal H}_{(A)}=0, \end{equation} \begin{equation} \label{c2} {\cal P}_k=\pi_{(\Gamma)}\partial_k{\rm ln}\Gamma+\partial_k\pi_{(\Gamma)}+ \Pi \partial_k\eta +2\nabla_l\pi^l_{(\omega) k}+\pi_{(A)}^lF_{lk}=0. \end{equation} \section{Interpretation of new variables} To treat the new variables $\Pi $ and $\eta $, we consider the flat-space limit \cite{Khved}: $\pi_{(\omega)}=\bar R=N^k=0$;\\$C(\eta)=1;\;S(\eta)=\eta,$ where the Hilbert action~(\ref{ham}) has the form $$ W_{GR}=\int dt d^3x\left\{\pi^k_{(A)}(\dot A_k-\partial_kA_0)- \Pi\dot \eta -N_c({\cal H}_{(A)}-\frac{\Pi}{\Gamma})\right\}. $$ After the reduction on the constraint shell $ {\cal H}_{EC} = 0$ in the gauge $N_c=1$, we get the expression for the conventional action of electrodynamics $$ W^{Red}=\int dt d^3x \left\{\pi^k_{(A)}(\dot A {}_k-\partial_kA_0)- {\cal H}_{(A)}\right\}, $$ Note that because \begin{equation} \label{em} \dot\eta=\frac{1}{\Gamma}. \end{equation} {\bf in this limit the variable $\eta$ can be treated as the time and the quantity $\Pi /\Gamma $ in the reduced phase space is a "reduced energy" like the the quantity $ \sqrt{p^2+m^2} $ is the spectral energy for a relativistic particle.} \section{The wave function of the Universe} Suppose that the constraint ${\cal H}_{EC}=0$ has the set of solutions $\Pi={\cal H}^{red}_{(\alpha)}\;\;\;,\alpha=1...m$. For each solution one can write down the corresponding reduced Hilbert action \begin{equation} \label{wred W^{red}_{GR(\alpha)}=\int dt \int d^3x \left(\sum_{\Phi=\omega , A,{\rm ln}\Gamma} \pi_{(\Phi)}\dot\Phi -{\cal H}^{red}_{(\alpha)}\dot\eta\right) +\bar W_S \end{equation} The quantization of the action leads to the Schr\"odinger type evolution equation \begin{equation} \label{wdw} \frac{1}{i}\frac{\delta}{\delta\eta}\Psi_{\alpha}=\hat{\cal H}{}^{red}_{(\alpha)} \Psi_{\alpha} . \end{equation} Let us consider the small time limit $(\eta)\sim 0 $ which corresponds to the small Universe $(a\sim S(\eta)\to0)$. According to equation~(\ref{cs}), in this region $S(\eta)\sim\eta$, $C(\eta)\sim 1 $, and the graviton term ${\pi }_{(\omega)}^2$ dominates in the energy density (\ref{c1}). The corresponding solutions of the constraints ${\cal H}_{EC}=0$ \begin{equation} \label{pi} \Pi={\cal H}^{red}_\pm = \pm\sqrt{6\pi_{(\omega)}^2}\frac{1}{\eta}. \end{equation} and the reduced action reads \begin{equation} \label{mis} W_{GR\pm} =\int^T_0dt\int_Vd^3x\left(\sum \pi_{(\omega)}\dot\omega \mp\sqrt{6\pi_{(\omega)}^2}\partial_0{\rm ln}\eta\right). \end{equation} One can verify that in the supposition of homogeneousity of the space \begin{equation} \label{homo ds^2=a^2[N^2_c(t)dt^2-A^2(r)dx^2];\;\;\; \;\;A(r)=\left(1+\frac{kr^2}{4r^2_0}\right)^{-1} \end{equation} \begin{equation} \label{fried} a(t)N_c(t)dt=dT_{Fried};\;\;\;\;k=0,\pm1, \end{equation} it follows from (\ref{mis}) the action for the Misner Universe \cite{Mis} (in details see~\cite{Ryan}): \begin{equation} \label{1mis} W_{GR\pm} = V_{(3)} \left(\pi_{(\omega)}(\omega(T)-\omega(0))\mp \sqrt{6\pi_{(\omega)}^2}{\rm ln}\frac{\eta(T)}{\eta(0)}\right), \end{equation} and the spectral decomposition for the wave function \begin{equation} \label{psq} \Psi = \int d\pi_{(\omega)} \left( A^+_{\pi_{(\omega)}}e^{iW_{GR+}} + A^+_{\pi_{(\omega)}}e^{iW_{GR-}} \right), \end{equation} where the role of the time is played by the logarithm of $\eta$, and $A^\pm$ are the operators of creation and annihilation of the Universe. Note that in the case of the homogeneous space the particular solutions of eq.~(\ref{cs}) read \begin{equation} \label{cos} C(\eta)=1,{\rm cos{\eta}},{\rm cosh{\eta}}; \quad S(\eta)= \eta,{\rm sin{\eta}},{\rm sinh{\eta}}, \quad \Gamma = r_0 \end{equation} respectively for \(k = 0, 1, -1\). The evolution of the Universe filled in by dust and radiation has been considered in \cite{Khved}. In this case the wave function of the Universe has the form $\Psi_\pm=\exp{\{\imath W_\pm^{red}(a)}\}$ with the reduced action $$ W^{red}_\pm=\pm V_{(3)} \int_{0}^{T}da \left[\underline{\pi}_{(a)}- \frac{1}{2}\frac{d}{da}(\underline{\pi}_{(a)}a)\right], $$ where $V_{(3)}=\int d^3xA^3(r)$ and $ \underline{\pi}_{(a)}=2\sqrt{\frac{3}{\kappa^2}}\left[{\cal H}_{(M)}- \frac{3ka^2}{r_0^2\kappa^2}\right]^{1/2} $ is the solution of the constraint (\ref{c1}) ${\cal H}_{EC}=0$. The phase of this function coincides (up to the energy factor) with the Friedmann time (\ref{fried}) for the dust case (${\cal H}_M=a\epsilon_{dust}$) \begin{equation} \label{d} W^{red}_{dust}=\frac{V_{(3)}\epsilon_{dust}}{2}T_{Fried}(a);\;\;\;\;\;\;\;\; T_{Fried}(a)=\int_{a(0)}^{a(T)}da\frac{6a}{\kappa^2 \underline{\pi}_{(a)}} \end{equation} and with the conformal time $N_cdt=\eta (t) r_0$ for the radiation (${\cal H}_M= \epsilon_{rad} $) \begin{equation} \label{r} W^{red}_{rad}=V_{(3)}\epsilon_{rad}\eta (a)r_0;\;\;\;\;\;\;\;\; \eta (a)r_0=\int_{a(0)}^{a(T)}da\frac{6}{\kappa^2 \underline{\pi}_{(a)}}. \end{equation} It is worth to note that the this clear correspondence between the quantum and classical physics ( the "phase time" and the "interval time") arises due to the maintenance of the time surface term $\frac{1}{2} \partial (\underline{\pi}_{(a)}a)$ in the Hilbert action. {\bf Thus, the time surface term helps us to establish the correspondence between the time as the phase of the ADM wave function of the Universe and the classical proper time as an invariant interval.} \section{Quantum scenario of the Weinberg-Salam Universe.} Let us consider the quantum scenario of the evolution of the Universe filled in by the Weinberg-Salam fields and described by the action \begin{equation} \label{ws} W=\int d^3xdt\sqrt{-g}\left\{-\frac{{}^{(4)}R}{2\kappa^2} +\frac{{}^{(4)}R \Phi^* \Phi }{6}+ \partial_{\mu}\Phi^* \partial^{\mu} \Phi -\gamma(\bar \Psi_{L}\Phi)\Psi_R+\cdots\right\} \end{equation} where $\Phi=\left(\begin{array}{c}\Phi_1\\ \Phi_2 \end{array}\right)$ is the doublet of the complex scalar fields; $\Psi_L$ and $\Psi_R$ are the left and right fermions. We keep only the term of the scalar-fermion interaction generating masses of the fermions \\ $m(\bar \Psi_{L2}\Psi_R+\bar \Psi_R\Psi_{L2})$ to show the evolution of the mass parameters with respect to the scale $a$. If we extract the scale factor \( a \) not only from the metric (\ref{metric}) $g_{\mu\nu}=a^2\tilde g_{\mu\nu}$, but also from all other matter fields $\Phi=a\varphi,\;\;\; \Psi=a^{-3/2}\Psi_c$, we can guarantee the classical limit of the massive fermion fields as the Friedmann dust of the Universe. The action in terms of the physical fields $\varphi,\Psi_c,\tilde g (\sqrt{-\tilde g}=N_c)$ has the form \begin{eqnarray W&=&\int d^4x\{N_c\left[-\frac{{}^{(4)}\tilde R}{2\kappa^2}\left(a^2- \frac{\kappa^2}{3}\varphi^*\varphi\right)-\frac{3}{\kappa^2}\partial_\nu a\partial^\nu a +\partial_\nu \varphi^*\partial^\nu\varphi-\gamma(\bar \Psi_{Lc}\varphi^)\Psi_{Rc}+\cdots\right]\nonumber\\ &&+\partial_\nu\left(N_c\left[\frac{3a^2}{\kappa^2}-\varphi^*\varphi \right]\frac{\partial^\nu a}{a}\right)\}\nonumber \end{eqnarray} Note that the configuration $a^2 =\frac{\kappa^2}{3}\varphi^*\varphi $ represents the singular point: in the vicinity of this point the sign before the four- dimensional curvature is changing. So, let us consider only the field configuration such that $[a^2-\frac{\kappa^2}{3}\varphi^*\varphi]=\rho^2 > 0$. For this case one can introduce new variables $$ a=\rho \cosh(\xi);\;\; \varphi_i=\sqrt{\frac{3}{\kappa^2}}\rho \sinh(\xi) n_i;\;\; n_1=\cos(\Theta)\exp{\{\imath \chi_1\}};\;\;n_2=\sin(\Theta)\exp{\{\imath \chi_2\}}, $$ where $\xi,\Theta,\chi_1,\chi_2$ are the angles of the scale-scalar mixing. For the homogeneous space~(\ref{homo}) with $V_{(3)}=1$ we get the action \begin{eqnarray} W&=&\int dt\{-\frac{D}{2N_c}\left(\frac{6}{\kappa^2}\right)+ N_c\left[\frac{k}{2r_0^2}\rho^2\left(\frac{6}{\kappa^2}\right) -\gamma\sqrt{\frac{3}{\kappa^2}}\rho\sinh(\xi)(\bar \Psi_{Lci}n_i)\Psi_{Rc}+\cdots\right] \nonumber\\ &&+\partial_0\left[ \frac{\rho^2}{2N_c}\left(\frac{\dot \rho}{\rho}+\dot \xi \tanh \xi\right) \right]\frac{6}{\kappa^2}\},\nonumber \end{eqnarray} where $D=-\dot \rho^2+\rho^2 \dot \xi^2+(\rho\sinh\xi)^2\left(\dot \Theta^2+ \sin^2(\Theta)\dot \chi^2_1+\cos^2(\Theta)\dot \chi^2_2\right)$ is the $SO(4,1)$-invariant differential form. By the Ostrogradsky method this action can be rewritten in terms of momenta $$ W=\int dt\{\sum_{\alpha=\xi,\Theta,\chi_1,\chi_2}\pi_{(\alpha)}\dot \alpha- \pi_{(\rho)}\dot \rho- N_c{\cal H}_{EC}+\frac{1}{2}\partial_0\left(\pi_{(\rho)}\rho\right)+ \frac{1}{2}\partial_0\left(\pi_{(\xi)}\tanh(\xi)\right)\} $$ $$ {\cal H}_{EC}=-\frac{1}{2}\frac{\kappa^2}{6}\pi^2_{(\rho)}-\frac{k\rho^2}{r^2_0} \frac{6}{\kappa^2} +\frac{{\cal K}^2}{2\rho^2}\frac{\kappa^2}{6} +\gamma\sqrt{\frac{3}{\kappa^2}}\rho\sinh(\xi)(\bar \Psi_{Lci}n_i)\Psi_{Rc}+\cdots $$ where ${\cal K}^2$ is the Kazimir operator of the $SO(4,1)$ group $$ {\cal K}^2=P^2_{(\xi)}+\frac{1}{\sinh^2(\xi)}\left(P^2_{(\Theta)}+ \frac{P^2_{\chi_1}}{\sin^2(\Theta)}+\frac{P^2_{\chi_2}}{\cos^2(\Theta)}\right). $$ We see that the variable $\rho$ plays the same role as the scale $a$ in the theory without scalar fields, and it is the time like variable. Note that our transition to new variables is similar to the Bekenstein transformation~\cite{Beken}. Repeating the canonical transformation (\ref{cantr}) $\pi_{(\rho)}\dot \rho- \frac{1}{2}\partial_0(\pi_{(\rho)}\rho)=\Pi\dot \eta$, where $\Gamma=r_0$, we get the expression for the action \begin{eqnarray} W&=&\int dt\{\sum_{\alpha}P_{(\alpha)}\dot \alpha-\Pi\dot \eta- N_c\left[ -\frac{\Pi }{r_0}+\frac{{\cal K}^2}{4\Pi r_0S^2(\eta)}+ \gamma\sqrt{\frac{3}{\kappa^2}}\rho\sinh(\xi)(\bar \Psi_{Lci}n_i)\Psi_{Rc} +\cdots\right]\nonumber\\ &&+\frac{1}{2}\partial_0\left(P_{(\xi)}\tanh (\xi)\right)\}\nonumber \end{eqnarray} This action describes the following ADM-scenario of the evolution of the Universe. In the small time limit $(\eta\sim 0)$ the Kazimir operator term dominates and the reduced system on the constraint ${\cal H}_{EC}=0$ has the form $$ W^{red}_{(\pm)}=\int_{T_s(0)}^{T_s(T)}dT_s\left[\sum_{\alpha} P_\alpha\frac{d\alpha}{dT_s}\mp{\cal K}(P_\alpha)+\frac{1}{2}\frac{d}{dT_S}(P_\xi\tanh(\xi)) \right];\;\;\;\;dT_s=\frac{d\eta}{2S(\eta)}. $$ The wave function of this system can be decomposed over the eigenfunctions of the Kazimir operator with the eigenvalues ${\cal K}_ \epsilon$ $$ \Psi(T_s=T_s(T)-T_s(0)|\alpha)=\sum\limits_{\epsilon} \left[A_\epsilon^{(+)}e^{+\imath {\cal K}_\epsilon T_s}\Psi_\epsilon(\alpha_T|\alpha_0) +A_\epsilon^{(-)}e^{-\imath {\cal K}_\epsilon T_s}\Psi_\epsilon^*(\alpha_T|\alpha_0)\right], $$ $$ \Psi_\epsilon(\alpha_T|\alpha_0)=Y_{SO(4,1)}(\alpha_T)Y^*_{SO(4,1)}(\alpha_0) \exp\left\{\imath P\xi\left(\tanh(\xi_T)-\tanh(\xi_0)\right)\right\}, $$ where $A^{(\pm)}$ are the operators of the creation and annihilation of the Universe, $Y_{SO(4,1)}$ is a unitary irreducible representation of the $SO(4,1)$ group. This wave function reproduces the physical picture of the Misner anisotropic Universe (\ref{mis}) discussed above in section 4. In the large time limit, the $SO(4,1)$ symmetry is broken, the Kazimir operator term disappears in comparison with the mass term. In this case, the masses of elementary particles in the Weinberg-Salam model are determined by the fixed values of angles of the scale-scalar field mixing and the ADM-observer gets the Friedmann cosmological models of radiation and dust, considered above. \vspace{0.3cm} \section{Conclusion} We have shown that including the time surface term in the canonical Hamiltonian formulation of GR helps us to extract the time-like variable and its conjugated momentum from the extended phase space. By taking into account the time surface term we represented here the new version of the Dirac - ADM Hamiltonian formalism for general relativity in the reduced phase space with the Schr\"odinger - like equation for a wave function describing the quantum evolution of the Universe. This evolution coincides with the Friedmann classical evolution of the dust filled Universe and shows that in GR like in special relativity there are two distinguished invariant time variables: {\bf the "phase time" of the ADM-observer (who constructs the Hamiltonian and measures the time as a phase of the wave function of the expanding Universe ) and the geometrical time of the Friedmann observer (who measures the time as an invariant proper interval and observes this expansion on the earth)}. In special relativity, the corresponding times are connected by the Lorentz transformation and they coincide only in the case when the rest frame of the Einstein observer coincides with the rest frame of a particle. Now the main question is to find the corresponding transformation from the rest frame of the ADM - observer to the Friedmann one. \vspace{0.3cm} {\bf Acknowledgment} \vspace{0.3cm} The authors thank Profs. A. Ashtekar, B M. Barbashov, G.T. Horowits, V.G. Kadyshevsky, K. Kuchar, D.A.Kirzhnitz, L.N. Lipatov, D.V.Volkov for useful discussions. One of the authors (V.N.P.) acknowledges the hospitality of the International Centre for Theoretical Physics in Trieste where this paper was finished. \vspace{0.5cm}
\section{Introduction} The origin of ${\cal CP}$ violation remains one of the outstanding problems in particle physics. In the attempt to understand this problem many experimental and theoretical efforts have been launched \cite{review}. One of the systems where it is possible to search for ${\cal CP}$ violation is the non-leptonic decay of hyperons. Although this has been known for many years \cite{marshak}, it is only recently that it has become conceivable to carry out an experimental program to look for ${\cal CP}$ violating signals in the decays of $\Xi$ and $\Lambda$ hyperons \cite{cernpro,proposal}. Of particular interest is the upcoming experiment E871 that expects to reach a sensitivity of $10^{-4}$ for the sum of asymmetries $A(\Lambda^0_-) + A(\Xi^-_-)$ \cite{proposal}. Unfortunately, the calculation of these asymmetries is plagued by theoretical uncertainties in the estimate of the hadronic matrix elements involved. Nevertheless, a conservative study of these asymmetries within the minimal standard model indicated that $A(\Lambda^0_-)$ is likely to occur at the level of a few times $10^{-5}$. In view of this, the potential results of E871 are very exciting. One of the questions we would like to answer is whether the phase in the CKM matrix of the three generation minimal standard model is the sole source of ${\cal CP}$ violation. The experimental information that we have so far is: \begin{itemize} \item A non-zero value of the parameter $\epsilon$ in kaon decays \cite{pdb}: \begin{equation} |\epsilon| = 2.26 \times 10^{-3} \end{equation} \item A measurement of the parameter $\epsilon^\prime$ \cite{epove}: \begin{equation} {\epsilon^\prime \over \epsilon} = \cases{ (2.3 \pm 0.65)\times 10^{-3} & NA31 \cr (0.74\pm0.52\pm0.29)\times 10^{-3} & E731 \cr} \label{eppexp} \end{equation} \end{itemize} The first result indicates that there is ${\cal CP}$ violation in nature, but it does not pinpoint its origin. The best one can say is that it is possible for the minimal standard model to accommodate this number. If the second number turns out to be non-zero it would establish the existence of direct $|\Delta S|=1$ ${\cal CP}$ violation, ruling out some superweak models. The current experimental numbers are consistent with the minimal standard model, although the theoretical calculations are also plagued with uncertainty from the evaluation of hadronic matrix elements. The present situation is, therefore, that there is no need for ${\cal CP}$ violation beyond the phase in the three generation CKM matrix\footnote{Except perhaps in the origin of the baryon asymmetry of the universe. We will not discuss that issue in this paper.}, but that other sources of ${\cal CP}$ violation have not been ruled out. The question we want to address in this paper is whether it is possible for E871 to find a non-zero asymmetry given its expected sensitivity and the current values of $\epsilon$ and $\epsilon^\prime/\epsilon$. To this end, and in keeping with the results of all the precise experiments conducted to date, we will assume that the minimal three-generation standard model is a very good low energy approximation to the electroweak interactions. We will, therefore, discuss any possible new physics in terms of an effective Lagrangian consistent with the symmetries of the standard model and will only look only at operators of dimension six. Our paper is organized as follows. In Section 2 we review the notation for ${\cal CP}$ violating observables in hyperon decays as well as the standard model estimate of $A(\Lambda^0_-)$. In section~3 we compute the contributions of ${\cal CP}$ violating four quark operators to $A(\Lambda^0_-)$ and the constraints that result from the measurements of ${\cal CP}$ violation in $K\rightarrow \pi\pi$. In section~4 we repeat this analysis for the two-quark operators of dimension six (so called penguin operators). Finally, we present our conclusions. \section{${\cal CP}$ Violation in $\Lambda^0 \rightarrow p \pi^-$} In this section we review the basic features of ${\cal CP}$ violation in the reaction $\Lambda^0 \rightarrow p \pi^-$, denoted by $(\Lambda^0_-)$. In the $\Lambda^0$ rest frame, $\vec{\omega}_{i,f}$ will denote unit vectors in the directions of the $\Lambda$ and $p$ polarizations, and $\vec{q}$ will denote the proton momentum. The isospin of the final state is $I= 1/2 {\rm ~or~} 3/2$, and each of these two states can be reached via a $\Delta I = 1/2 {\rm ~or~} 3/2$ weak transition respectively. There are also two possibilities for the parity of the final state. They are the $s$-wave, $l=0$, parity odd state (thus reached via a parity violating amplitude); and the $p$-wave, $l=1$, parity even state reached via a parity conserving amplitude. A model independent analysis of the decay can be done by writing the most general matrix element consistent with Lorentz invariance: \cite{marshak} \begin{equation} {\cal M} = G_F m_\pi^2 \overline{u}_P(A-B\gamma_5)u_\Lambda . \label{gmatel} \end{equation} It is customary to introduce the quantities: \begin{eqnarray} s&=&A\nonumber \\ p&=&\biggl({|\vec{q}|\over E_P + M_P}\biggr) B \end{eqnarray} to write the total decay: \begin{equation} \Gamma = {|\vec{q}|(E_P+M_P) \over 4 \pi M_\Lambda}G_F^2 m_\pi^4 \left(|s|^2+|p|^2\right). \label{drate} \end{equation} The angular distribution is proportional to: \begin{equation} {d\Gamma \over d\Omega} \sim 1+ \gamma\vec{\omega}_i\cdot\vec{\omega}_f + (1-\gamma)\hat{q}\cdot\vec{\omega}_i \hat{q}\cdot\vec{\omega}_f + \alpha \hat{q}\cdot(\vec{\omega}_i+\vec{\omega}_f) + \beta \hat{q}\cdot(\vec{\omega}_f\times\vec{\omega}_i), \label{msqsim} \end{equation} where we have used the standard notation \cite{marshak}: \begin{equation} \alpha \equiv {2 {\rm Re} s^*p \over |s|^2 + |p|^2}, \;\; \beta \equiv {2 {\rm Im} s^*p \over |s|^2 + |p|^2}, \;\; \gamma \equiv {|s|^2 - |p|^2 \over |s|^2 + |p|^2} . \label{albega} \end{equation} If the proton polarization is not observed, $\alpha$ is the parameter that governs the angular distribution: \begin{equation} {d\Gamma \over d\Omega}={\Gamma \over 4 \pi} \left(1+\alpha \hat{q}\cdot \vec{\omega}_i\right). \label{alonly} \end{equation} Similarly, if the initial $\Lambda$ is unpolarized, $\alpha$ determines the polarization of the proton: \begin{equation} \vec{\cal P}_p = \alpha_\Lambda \hat{q} \label{otheral} \end{equation} E871 will not measure the correlations governed by the parameter $\beta$ so we will not deal with it in this paper. The ${\cal CP}$-odd observable $A(\Lambda^0_-)$ is constructed by comparing the parameter $\alpha$ in the reaction $\Lambda^0 \rightarrow p \pi^-$ with the corresponding parameter $\overline{\alpha}$ in the reaction $\overline{\Lambda}^0 \rightarrow \overline{p} \pi^+$. One can show that ${\cal CP}$ symmetry predicts that: \begin{equation} \overline{\alpha} = - \alpha \label{cppred} \end{equation} so that a ${\cal CP}$ odd observable is: \cite{donpa} \begin{equation} A \equiv {\alpha \Gamma + \overline{\alpha}\overline{\Gamma} \over \alpha \Gamma - \overline{\alpha}\overline{\Gamma}} \approx {\alpha + \overline{\alpha} \over \alpha - \overline{\alpha}} \label{cpobs} \end{equation} Other possible ${\cal CP}$ odd observables have been discussed in the literature: a rate asymmetry that is significantly smaller than $A$ \cite{donpa}; and an asymmetry based on the parameter $\beta$ that won't be accessible to E871. For these reasons we concern ourselves with the observable $A(\Lambda^0_-)$. \footnote{In fact E871 will be sensitive to the sum $A(\Lambda^0_-) + A(\Xi^-_-)$. An analysis of $A(\Xi^-_-)$ parallels the one we will carry out, but doesn't really affect our conclusions given the inherent uncertainties in the computation of matrix elements. It has also been argued that $A(\Xi^-_-)$ is probably smaller than $A(\Lambda^0_-)$ due to much smaller strong rescattering phases \cite{lusawi}.} It is convenient to decompose the amplitudes according to isospin, and to introduce the following notation for the phases: \begin{eqnarray} s(\Lambda^0_-)= - \sqrt{2/3}s_{1}e^{i(\delta^s_1+\phi^s_1)}+\sqrt{1/3} s_{3}e^{i(\delta^s_3+\phi^s_3)}\nonumber \\ p(\Lambda^0_-)= - \sqrt{2/3}p_{1}e^{i(\delta^p_1+\phi^p_1)}+\sqrt{1/3} p_{3}e^{i(\delta^p_3+\phi^p_3)} \label{overnot} \end{eqnarray} where $\delta^I_J$ is the strong rescattering phase for the pion nucleon system and $\phi^I_J$ is the ${\cal CP}$ violating phase. In terms of these quantities one finds: \cite{donpa} \begin{eqnarray} A(\Lambda^0_-)&=& -\tan\left(\delta^p_1-\delta^s_1\right) \sin\left(\phi^p_1-\phi^s_1\right)\biggl[ 1 \nonumber \\ &+&{1\over \sqrt{2}}{s_3 \over s_1}\biggl( {\cos(\delta^p_1-\delta^s_3) \over \cos (\delta^p_1-\delta^s_1)}-{\sin(\delta^p_1-\delta^s_3) \over \sin(\delta^p_1-\delta^s_1)}{\sin(\phi^p_1-\phi^s_3) \over \sin(\phi^p_1-\phi^s_1)}\biggr) \nonumber \\ &+&{1\over \sqrt{2}}{p_3 \over p_1}\biggl( {\cos(\delta^p_3-\delta^s_1) \over \cos (\delta^p_1-\delta^s_1)}-{\sin(\delta^p_3-\delta^s_1) \over \sin(\delta^p_1-\delta^s_1)}{\sin(\phi^p_3-\phi^s_1) \over \sin(\phi^p_1-\phi^s_1)}\biggr) \biggr] \label{approxas} \end{eqnarray} Experimentally we know the values of: \begin{itemize} \item the strong rescattering phases \cite{roper}: \begin{equation} \delta_1 \approx 6.0^\circ,\;\;\delta_3 \approx -3.8^\circ,\;\; \delta_{11}\approx -1.1^\circ,\;\;\delta_{31}\approx-0.7^\circ \label{stphex} \end{equation} with all the errors on the order of $1^\circ$. \item the $\Delta I=3/2$ amplitudes are much smaller than the $\Delta I=1/2$ amplitudes \cite{pdbee}: \begin{equation} s_{3}/s_{1} = 0.027 \pm 0.008, \;\; p_{3}/p_{1}=0.03\pm 0.037 \label{expamp} \end{equation} \item the $s$ and $p$ amplitudes (assuming they are dominated by the ${\cal CP}$ conserving, $\Delta I =1/2$, transitions): \begin{eqnarray} s \approx -\sqrt{2 \over 3} s_1 &=& 1.47 \pm 0.01 \nonumber \\ p \approx -\sqrt{2 \over 3} p_1 &=& \biggl({|\vec{q}|\over E_P + M_P}\biggr) (9.98 \pm 0.24) \end{eqnarray} \end{itemize} Substituting the experimental numbers for the amplitudes and strong rescattering phases one gets: \begin{equation} A(\Lambda^0_-)\approx 0.13 \sin(\phi^p_1-\phi^s_1) +0.001 \sin(\phi^p_1-\phi^s_3) -0.0024 \sin(\phi^p_3-\phi^s_1) \label{alambda} \end{equation} \subsection{Standard model calculation} In the case of the minimal standard model, the ${\cal CP}$ violating phase resides in the CKM matrix. For low energy transitions, this phase shows up as the imaginary part of the Wilson coefficients in the effective weak Hamiltonian. In the notation of Buras \cite{buras}, \begin{equation} H_W^{SM} = {G_F \over \sqrt{2}}V^*_{ud}V_{us}\sum_i c_i(\mu)Q_i(\mu) + {\rm ~h.~c.} \label{effweak} \end{equation} $Q_i(\mu)$ are four quark operators, and $c_i(\mu)$ are the Wilson coefficients that are usually written as: \begin{eqnarray} c_i(\mu) &=&z_i(\mu)+\tau y_i(\mu) \nonumber \\ \tau &=& - {V^*_{td}V_{ts} \over V^*_{ud}V_{us}} \end{eqnarray} with the ${\cal CP}$ violating phase being the phase of $\tau$. Numerical values for these coefficients can be found, for example, in Buchalla {\it et. al.} \cite{buras}. The calculation of the weak phases would proceed by evaluating the hadronic matrix elements of the four-quark operators in Eq.~\ref{effweak} to obtain real and imaginary parts for the amplitudes, schematically: \begin{equation} \matel{p \pi}{H_w^{eff}}{\Lambda^0}|^I_\ell = {\rm Re}M^I_\ell + i {\rm Im} M^I_\ell , \label{schematic} \end{equation} and to the extent that the ${\cal CP}$ violating phases are small, they can be approximated by \begin{equation} \phi^I_\ell \approx { {\rm Im}M^I_\ell \over {\rm Re}M^I_\ell}. \label{smallph} \end{equation} At present, however, we do not know how to compute the matrix elements so we cannot actually implement this calculation. For a simple estimate, we can take the real part of the matrix elements from experiment (assuming that the measured amplitudes are real, that is, that ${\cal CP}$ violation is small), and compute the imaginary parts in vacuum saturation. This approach provides a conservative estimate for the weak phases because the model calculation of the real part of the amplitudes is smaller than the experimental value. Nevertheless, the numbers should be viewed with great caution. The approximate weak phases estimated in vacuum saturation are: \cite{steger} \begin{eqnarray} \phi^1_s &\approx& -3 y_6 {\rm Im}\tau \nonumber \\ \phi^1_p &\approx& -0.3 y_6 {\rm Im}\tau \nonumber \\ \phi^3_s &\approx& \left[3.6(y_1+y_2)+2.7(y_7+3y_8) {B_0^2 \over m_K^2}\right] {\rm Im}\tau \nonumber \\ \phi^3_p &\approx& \left[0.5(y_1+y_2)-0.4(y_7+3y_8) {B_0^2 \over m_K^2}\right] {\rm Im}\tau \label{approxph} \end{eqnarray} These provide numerical estimates using the values for the Wilson coefficients\footnote{For $\mu=1$~GeV, $\Lambda_{QCD}=200$~MeV} of Buchalla {\it et. al.} \cite{buras}, $y_6 \approx -0.08$; and the value of $B_0$ given in the appendix. For the quantity ${\rm Im}\tau$ (we use the Wolfenstein parameterization of the CKM matrix) we take: \begin{equation} {\rm Im}\tau = A^2 \lambda^4 \eta \leq 0.001 \label{imtau} \end{equation} Putting all the numbers together, and using the upper limit in Eq.~\ref{imtau} yields:\footnote{See Ref.~\cite{steger} for additional discussions of this calculation.} \begin{equation} A(\Lambda^0_-) \approx 3 \times 10^{-5} \label{vsnumres} \end{equation} Other models of ${\cal CP}$ violation contain additional short distance operators with ${\cal CP}$ violating phases \cite{wein,moha,senj} and predict different values for $A(\Lambda^0_-)$ \cite{donpa,chang}. A summary of results can be found, for example, in Ref.~\cite{hedpf}. \section{Four-quark Operators} We now study, in a model independent manner, the contributions to $A(\Lambda^0_-)$ that occur due to physics beyond the minimal standard model. In this section we look at the effect of all the four-quark operators and in the next section we discuss the two-quark operators. We assume that the physics that lies beyond the minimal standard model is characterized by a scale $\Lambda \gg M_W$ and, therefore, that its most important low energy effects can be parameterized by the lowest dimension operators of the most general effective Lagrangian consistent with the symmetries of the standard model. Such a Lagrangian has been written down by Buchm\"{u}ller and Wyler \cite{buchmuller}. In the appendix we list all the operators that occur at dimension six with $|\Delta S| =1$. The calculation then proceeds as in the previous section, but with the effective Hamiltonian \begin{equation} H_{eff} = H_W^{SM} + {g^2 \over \Lambda^2} \biggl(\sum_i\lambda_i {\cal O}^{new}_i + {\rm ~h.~c.}\biggr) \label{effh} \end{equation} To the usual, QCD corrected, standard weak Hamiltonian of the previous section we add all the four-fermion operators with $|\Delta S|=1$ that come from the new physics sector. We will sidestep the issue of the possible origin of the effective ${\cal CP}$ violating operators. We use the notation of Ref.~\cite{buchmuller} as detailed in the Appendix. These operators violate ${\cal CP}$ if the coupling $\lambda_i$ has an imaginary part. The normalization has been chosen for convenience. \subsection{$K_L \rightarrow \pi \pi$ and $\epsilon^\prime / \epsilon$} The standard notation for the $K \rightarrow \pi \pi$ amplitudes is: \begin{eqnarray} A(K^0 \rightarrow \pi^+ \pi^-) &=& A_0 e^{i\delta_0}+{A_2 \over \sqrt{2}} e^{i\delta_2} \nonumber \\ A(K^0 \rightarrow \pi^0 \pi^0)&=& A_0 e^{i\delta_0}- \sqrt{2}A_2 e^{i\delta_2} \end{eqnarray} where $\delta_{0,2}$ are the strong $\pi\pi$ scattering phases in the $I=0,2$ channel. The amplitudes $A_0$ and $A_2$ are real unless there is ${\cal CP}$ violation. Experimentally it is known that the $\Delta I = 3/2$ amplitude $A_2$ is much smaller than the $\Delta I = 1/2$ amplitude $A_0$: \begin{equation} \omega \equiv {{\rm Re}A_2 \over {\rm Re}A_0} \approx {1 \over 22} \end{equation} The contribution of the dominant penguin operator (${\cal O}_6$ in the notation of Ref.~\cite{buras}) to $\epsilon^\prime/\epsilon$ is given by: \begin{equation} \biggl({\epsilon^\prime \over \epsilon }\biggr)_6 = -{\omega \over \sqrt{2} |\epsilon|} {{\rm Im}(A_0)_6 \over |A_0|}= {\omega \over 2 |\epsilon|}{G_F \over |A_0|} y_6 \lambda {\rm Im}\tau <\pi^+\pi^-|{\cal O}_6|K^0> \end{equation} The hadronic uncertainty enters the calculation through the matrix element of the four-quark operator. We will use the estimate of Ref.~\cite{buras} for the matrix element of ${\cal O}_6$: \begin{equation} <\pi^+\pi^-|{\cal O}_6|K^0>\bigg|_{I=0} = -4 \sqrt{2}f_\pi {m_K^2 - m_\pi^2 \over \Lambda_\chi^2}\biggl({m_K^2 \over m_s + m_d}\biggr)^2 \approx -0.26 {\rm GeV}^3 \end{equation} Using the values $A=0.9$, $\lambda =0.22$ and $\eta =0.5$ one finds that: \begin{equation} \biggl({\epsilon^\prime \over \epsilon}\biggr)_6\approx 1.5 \times 10^{-3} \end{equation} The usual standard model analysis of $\epsilon^\prime/\epsilon$ consists of computing this contribution of the ``penguin'' operator, and of normalizing all other contributions to it in terms of a parameter $\Omega$: \begin{equation} {\epsilon^\prime \over \epsilon}=\biggl({\epsilon^\prime \over \epsilon} \biggr)_6 \biggl( 1 -\Omega_{SM}-\Omega_{NEW}\biggr) \end{equation} $\Omega_{SM}$ is given, for example, in Ref.~\cite{buras}, and we have introduced an analogous term $\Omega_{NEW}$ for the contributions of the new four-quark operators. Given the experimental result in Eq.~\ref{eppexp}, we will place bounds on the new physics by requiring, conservatively, that $\Omega_{NEW}\leq 1$. We find: \begin{eqnarray} \Omega_{NEW} &=& 8\biggl({M_W \over \Lambda}\biggr)^2\sum_i \biggl({ {\rm Im}\lambda_i \over A^2 \lambda^5 \eta}\biggr)\biggl[ {<\pi^+\pi^-|{\cal O}_i|K^0>_{I=0}\over y_6 <\pi^+\pi^-|{\cal O}_6|K^0>_{I=0} } \nonumber \\ && - {\sqrt{2}\over \omega} {<\pi^+\pi^-|{\cal O}_i|K^0>_{I=2}\over y_6 <\pi^+\pi^-|{\cal O}_6|K^0>_{I=0} } \biggr] \label{epprime} \end{eqnarray} Because there is no way at present to compute the matrix elements of four-quark operators reliably, we will simply use vacuum saturation. The new contributions to $\epsilon^\prime$ can thus be computed with the aid of the matrix elements listed in Table~\ref{t: matelkpp}. We use, as before, $A^2\lambda^5\eta \approx 2 \times 10^{-4}$ and we explicitly separate the contributions from the different isospin components of each operator for later convenience. We thus write: \begin{equation} \Omega_{NEW} = 4 \times 10^4 \biggl({M_W \over \Lambda}\biggr)^2\sum_i {\rm Im}\lambda_i \biggl( \omega_{0i} + \omega_{2i} \biggr) \label{numcoe} \end{equation} where $\omega_{0,2i}$ refers to the $\Delta I=1/2,3/2$ component of ${\cal O}_i$. We present numerical results for $\omega_{0,2i}$ in Table~\ref{t: omega}. \begin{table}[tbh] \centering \caption[]{Numerical coefficients for Eq.~\ref{numcoe}} \begin{tabular}{|c|c|c|} \hline Operator & $\omega_0$ & $\omega_2$ \\ \hline ${\cal O}_{qq}^{(1,1)}$ & $-0.06$ & 0 \\ ${\cal O}_{qq}^{(8,1)}$ & $-0.3$ & 0 \\ ${\cal O}_{qq}^{(1,3)}$ & $-0.3$ & 0 \\ ${\cal O}_{qq}^{(8,3)}$ & $0.32$ & 0 \\ \hline ${\cal O}_{dd}^{(1)}$ & $0.08$ & $2.5$ \\ ${\cal O}_{ud}^{(1)}$ & $-0.02$ & $-2.5$ \\ ${\cal O}_{dd}^{(8)}$ & $0.1$ & $3.3$ \\ ${\cal O}_{ud}^{(8)}$ & $0.2$ & $-3.3$ \\ \hline ${\cal O}_{qu}^{(1)}$ & $2.4$ & $-36.8$ \\ ${\cal O}_{qu}^{(8)}$ & $0.1$ & $3.3$ \\ ${\cal O}_{qd}^{(1)}$ & $1.5$ & 0 \\ ${\cal O}_{qsd}^{(1)}$ & $-3.9$ & $36.8$ \\ ${\cal O}_{qsd}^{(8)}$ & $-0.1$ & $-3.3$ \\ \hline ${\cal O}_{qsq}^{(1)}$ & $1.8$ & $-31.2$ \\ ${\cal O}_{qsq}^{(8)}$ & $-3.5$ &$33$ \\ ${\cal O}_{qqs}^{(1)}$ & $-3.5$ & $31$ \\ ${\cal O}_{qqs}^{(8)}$ & $2.1$ & $-33$ \\ ${\cal O}_{qq}^{(1s)}$ & $1.8$ & 0 \\ ${\cal O}_{qq}^{(8s)}$ & $1.3$ & 0 \\ \hline \end{tabular} \label{t: omega} \end{table} Requiring that $\Omega_{NEW} < 1$, we can constrain the size of the ${\cal CP}$ violating couplings Im~$\lambda_i/\Lambda^2$. By assuming that there is no accidental cancellation between the contributions of different operators to $\Omega_{NEW}$ we may constrain each operator separately. The isospin decomposition is useful because it is possible to construct combinations of operators with definite isospin transformation properties. The constraints that apply to operators that are purely $\Delta I =1/2$ are different from those that apply to operators that are purely $\Delta I = 3/2$. \subsection{$K^0-\overline{K}^0$ Mixing and $\epsilon$} In general, $\epsilon^\prime$ provides tighter constraints on new ${\cal CP}$ violating interactions that does $\epsilon$. Nevertheless, it is necessary to consider constraints from $\epsilon$ because the ones that arise from $\epsilon^\prime$ do not apply to parity conserving operators that do not contribute to the decay $K^0 \rightarrow \pi \pi$. In the operator basis that we are using, all the operators have parity conserving and violating components. However, it is possible to construct parity conserving combinations of operators just as it is possible to construct combinations of operators with definite isospin. All of the $|\Delta S|=1$ four-quark operators that we consider contribute to $\epsilon$ when combined with a second $|\Delta S|=1$ vertex from the usual weak Hamiltonian. In terms of the $K^0-\overline{K}^0$ mixing matrix, each operator gives a contribution to $\epsilon$ of the form: \begin{equation} |\epsilon|_i \approx {1 \over \sqrt{2}}{|{\rm Im}M_{12}|_i \over \Delta m_k} \end{equation} We estimate the long distance contributions to Im$M_{12}$ due to intermediate pion and eta poles \cite{dohowe}. Using the matrix elements of Table~\ref{t: matelep} we find that there is a cancellation between the contributions of the pion and octet-eta poles at leading order in $SU(3)$ breaking. This situation is unfortunate because it makes the estimates very unreliable. For our purposes we will use the model of Ref.~\cite{dhlin} to deal with this problem. The contribution of each operator to $\epsilon$ is given by: \begin{eqnarray} |\epsilon|_{{\cal O}_i} & = & \sqrt{2} g^2 {g_8 \over M_W^2}\biggl( {m_K \over \Delta m_K}\biggr) |{\rm Im} \lambda_i| \biggl({M_W \over \Lambda}\biggr)^2{m_K^2 \over m_K^2-m_\pi^2} |\xi_i|\nonumber \\ &\approx& 9.3 \ |{\rm Im} \lambda_i| \biggl({M_W \over \Lambda}\biggr)^2|\xi_i| \end{eqnarray} where $g_8$ is defined in Eq.~\ref{smkpipole}, and $\xi_i$ is given according to the model of Ref.~\cite{dhlin} by: \begin{eqnarray} \xi_i &=& {1\over \sqrt{2}f_\pi^2 m_K^2} {f_K \over f_\pi} \biggl\{<\pi^0|{\cal O}_i|\overline{K}^0> + {<\eta_8|{\cal O}_i|\overline{K}^0> \over \sqrt{3} } \nonumber \\ &&\cdot \biggl\{\biggl({m_K^2 - m_\pi^2 \over m_K^2 - m_\eta^2}\biggr) \biggl[(1+\xi)\cos\theta+2\sqrt{2}\rho\sin\theta\biggr] \biggl[{f_{\eta_8}\over f_\pi}\cos\theta-\sqrt{2} {f_{\eta_0} \over f_\pi}\sin\theta\biggr] \nonumber \\ &&+ \biggl({m_K^2 - m_\pi^2 \over m_K^2 - m_\eta^{\prime 2}}\biggr) \biggl[(1+\xi)\sin\theta-2\sqrt{2}\rho\cos\theta\biggr] \biggl[{f_{\eta_8}\over f_\pi}\sin\theta+\sqrt{2} {f_{\eta_0} \over f_\pi}\cos\theta\biggr] \biggr\}\biggr\} \label{newep} \end{eqnarray} We choose the parameters that Ref.~\cite{dhlin} considers more physical: $\theta = -20^\circ$, $\xi = 0.17$, $f_{\eta 8} = 1.25 f_\pi$, $f_{\eta 0}=1.04 f_\pi$. Once more we present separate results for the $\Delta I = 1/2, 3/2$ components of each operator in Table~\ref{t: xi}. We emphasize again that we present our results in this form because it is possible to construct combinations of operators that have definite isospin transformation properties. For the $\Delta I = 1/2$ component, there is sensitivity to the parameters in the model of Ref.~\cite{dhlin}. We illustrate this by presenting results for $\rho=0.8$, $\rho = 1.2$, and for just the pion pole. For the $\Delta I= 3/2$ component there is only a pion pole. \begin{table}[tbh] \centering \caption[]{Factors $\xi_i$ for Eq.~\ref{newep}.} \begin{tabular}{|c|c|c|c|c|} \hline Operator & $\xi_{i,1/2} (\rho =0.8)$ & $\xi_{i,1/2} (\rho =1.2)$ & $\xi_{i,1/2}$($\pi$-only) & $\xi_{i,3/2}$ \\ \hline ${\cal O}_{qq}^{(1,1)}$ & $ -0.24 $ & $0.14$ & $-0.04$ & $0$ \\ ${\cal O}_{qq}^{(8,1)}$ & $-0.41 $ & $-0.12$ & $-0.22$ & $0$ \\ ${\cal O}_{qq}^{(1,3)}$ & $-0.33 $ & $-0.17$ & $-0.21$ & $0$ \\ ${\cal O}_{qq}^{(8,3)}$ & $-0.16 $ & $0.71$ & $0.22$ & $0$ \\ \hline ${\cal O}_{dd}^{(1)}$ & $-0.18 $ & $0.04$ & $-0.06$ & $-0.11$ \\ ${\cal O}_{ud}^{(1)}$ & $-0.06 $ & $0.1$ & $0.01$ & $0.11$ \\ ${\cal O}_{dd}^{(8)}$ & $-0.23 $ & $0.06$ & $-0.07$ & $-0.15$ \\ ${\cal O}_{ud}^{(8)}$ & $-0.18 $ & $-0.18$ & $-0.15$ & $0.15$ \\ \hline ${\cal O}_{qu}^{(1)}$ & $-1.8 $ & $-1.7$ & $-1.5$ & $1.5$ \\ ${\cal O}_{qu}^{(8)}$ & $-0.05 $ & $0.24$ & $0.07$ & $0.15$ \\ ${\cal O}_{qd}^{(1)}$ & $-4.2 $ & $-1.2$ & $-2.2$ & $0$ \\ ${\cal O}_{qsd}^{(1)}$ & $-2.4 $ & $0.57$ & $-0.75$ & $-1.5$ \\ ${\cal O}_{qsd}^{(8)}$ & $-0.23 $ & $0.06$ & $-0.07$ & $-0.15$ \\ \hline ${\cal O}_{qsq}^{(1)}$ & $-1.7 $ & $-2.2$ & $-1.6$ & $1.2$ \\ ${\cal O}_{qsq}^{(8)}$ & $0.46 $ & $-2.1$ & $-0.65$ & $-1.3$ \\ ${\cal O}_{qqs}^{(1)}$ & $0.22 $ & $-2.6$ & $-1.0$ & $-1.2$ \\ ${\cal O}_{qqs}^{(8)}$ & $-1.6 $ & $-1.6$ & $-1.3$ & $1.3$ \\ ${\cal O}_{qq}^{(1s)}$ & $-1.5 $ & $-4.8$ & $-2.6$ & $0$ \\ ${\cal O}_{qq}^{(8s)}$ & $-1.1$ & $-3.7$ & $-2.0$ & $0$ \\ \hline \end{tabular} \label{t: xi} \end{table} \subsection{$\Lambda \rightarrow p \pi^-$ and $A(\Lambda^0_-)$} The starting point of the calculation is Eq.~\ref{alambda}. We study the effect of the new physics one operator at a time and always assume that the ${\cal CP}$ violating amplitudes are small, so that the experimental value of the amplitudes is approximately equal to the ${\cal CP}$ conserving amplitude. All the ${\cal CP}$ violating phases are then small and we can write: \begin{eqnarray} A(\Lambda^0_-) &\approx & 3 \times 10^{-5} \nonumber \\ &+& \sum_i \biggl( 0.13 (\phi^p_1-\phi^s_1) +0.001 (\phi^p_1-\phi^s_3) -0.0024 (\phi^p_3-\phi^s_1)\biggr)_i \label{anew} \end{eqnarray} where the sum runs over all the operators in Eq.~\ref{effh}. We carry out the calculation in the same manner as the standard model analysis of the previous section \cite{steger}. That is, we compute the imaginary part of the amplitudes by taking matrix elements of each new four-quark operator in vacuum saturation. Further, we will not compute perturbative QCD corrections to the effective Hamiltonian of the new physics sector. We will also assume that the new physics does not significantly alter the ${\cal CP}$ conserving amplitudes, but we will comment on this later on. As discussed in Ref.~\cite{steger}, this vacuum saturation calculation is not reliable at all, nevertheless, we will use it for lack of anything better. Calculating the imaginary part of the amplitudes taking the real part from experiment as in the previous section, we find that each operator ${\cal O}_i$ induces the following phases: \begin{eqnarray} \biggl(\phi^p_1\biggr)_i &=& -8 {G_F \over \sqrt{2}} \biggl({M_W \over \Lambda}\biggr)^2 {\rm Im}\lambda_i {<p \pi^-|{\cal O}_i|\Lambda>^P_1 \over 9.98 G_F m_\pi^2 } \nonumber \\ \biggl(\phi^s_1\biggr)_i &=& 8 {G_F \over \sqrt{2}} \biggl({M_W \over \Lambda}\biggr)^2 {\rm Im}\lambda_i {<p \pi^-|{\cal O}_i|\Lambda>^S_1 \over 1.47G_F m_\pi^2 } \nonumber \\ \biggl(\phi^p_3\biggr)_i &=& 8 {G_F \over \sqrt{2}} \biggl({M_W \over \Lambda}\biggr)^2 {\rm Im}\lambda_i {<p \pi^-|{\cal O}_i|\Lambda>^P_3 \over 0.21 G_F m_\pi^2 } \nonumber \\ \biggl(\phi^s_3\biggr)_i &=& -8 {G_F \over \sqrt{2}} \biggl({M_W \over \Lambda}\biggr)^2 {\rm Im}\lambda_i {<p \pi^-|{\cal O}_i|\Lambda>^S_3 \over 0.03 G_F m_\pi^2} \nonumber \\ \label{hypphase} \end{eqnarray} The matrix elements are estimated in vacuum saturation and listed in Table~\ref{t: matel} in the appendix. Numerically we find: \begin{equation} A(\Lambda^0_-) \approx 3 \times 10^{-5} + \biggl({M_W \over \Lambda}\biggr)^2 \sum_i {\rm Im}\lambda_i a_i \label{newal} \end{equation} where the coefficients $a_i$ are listed in Table~\ref{t: hyp}. We present two different values: in the first column we include only the $\Delta I = 1/2$ component of each operator, whereas in the second column we include both isospin components. We can see from Table~\ref{t: hyp} that the ${\cal CP}$ violating asymmetry $A(\Lambda^0_-)$ is dominated by the interference of the $s$ and $p$ waves in the $\Delta I =1/2$ amplitude, as can be anticipated from Eq.~\ref{alambda}. We also see from Table~\ref{t: hyp} that $a_i$ is of order one in some cases. Eq.~\ref{newal} then tells us that a measurement of $A(\Lambda^0_-)$ at the $10^{-4}$ level is sensitive, in principle, to new ${\cal CP}$ violating interactions generated at a scale $\Lambda \leq 8$~TeV and is thus potentially interesting. \begin{table}[tbh] \centering \caption[]{Factors $a_i$ for $A(\Lambda^0_-)_{NEW}$ in Eq.~\ref{newal}.} \begin{tabular}{|c|c|c|} \hline Operator & $(a_i)_{1/2}$ & $a_i$ \\ \hline ${\cal O}_{qq}^{(1,1)}$ & $0.03$ & $0.03$ \\ ${\cal O}_{qq}^{(8,1)}$ & $0.15$ & $0.14$ \\ ${\cal O}_{qq}^{(1,3)}$ & $0.14$ & $0.14$ \\ ${\cal O}_{qq}^{(8,3)}$ & $-0.15$ & $-0.15$ \\ \hline ${\cal O}_{dd}^{(1)}$ & $-0.05$ & $-0.06$ \\ ${\cal O}_{ud}^{(1)}$ & $0.01$ & $0.02$ \\ ${\cal O}_{dd}^{(8)}$ & $-0.06$ & $-0.08$ \\ ${\cal O}_{ud}^{(8)}$ & $-0.1$ & $-0.1$ \\ \hline ${\cal O}_{qu}^{(1)}$ & $-1.3$ & $-1.1$ \\ ${\cal O}_{qu}^{(8)}$ & $-0.05$ & $-0.08$ \\ ${\cal O}_{qd}^{(1)}$ & $1.5$ & $1.5$ \\ ${\cal O}_{qsd}^{(1)}$ & $-0.6$ & $-0.8$ \\ ${\cal O}_{qsd}^{(8)}$ & $0.05$ & $0.08$ \\ \hline ${\cal O}_{qsq}^{(1)}$ & $-1.4$ & $-1.2$ \\ ${\cal O}_{qsq}^{(8)}$ & $0.6$ & $0.7$ \\ ${\cal O}_{qqs}^{(1)}$ & $-0.9$ & $-1.0$ \\ ${\cal O}_{qqs}^{(8)}$ & $-1.1$ & $-1.0$ \\ ${\cal O}_{qq}^{(1s)}$ & $1.8$ & $1.8$ \\ ${\cal O}_{qq}^{(8s)}$ & $1.4$ & $1.3$ \\ \hline \end{tabular} \label{t: hyp} \end{table} We can use the constraints from ${\cal CP}$ violation in kaon decays to place bounds on the magnitude of $A(\Lambda^0_-)$ that each of the four quark operators can induce. In general, the bounds coming from direct ${\cal CP}$ violation in $\epsilon^\prime$ are stronger than those coming from $\epsilon$. However, it is necessary to consider both because it is possible to construct parity conserving combinations of operators that do not contribute to $K\rightarrow \pi \pi$ amplitudes and, thus, evade the bounds from $\epsilon^\prime$. Similarly, $\epsilon^\prime$ places stronger constraints on $\Delta I =3/2$ operators than on $\Delta I =1/2$ operators due to the enhancement factor of $1/\omega$ in Eq.~\ref{epprime}. To take into account these distinctions, we list in Table~\ref{t: result} the bounds on each of the weak phases separately. The blank entries indicate that there is no bound because the particular operator does not contribute to that amplitude. \begin{table}[tbh] \centering \caption[]{Bounds on the phases that enter $A(\Lambda^0_-)$.} \begin{tabular}{|c|c|c|c|c|} \hline Operator & $\phi^p_1 \times 10^{5}$ & $\phi^s_1\times 10^{5}$ & $\phi^p_3\times 10^{5}$ & $\phi^s_3\times 10^{5}$\\ \hline ${\cal O}_{qq}^{(1,1)}$ & $2.9$ & $-10$ & $--$ & $--$ \\ ${\cal O}_{qq}^{(8,1)}$ & $9.0$ & $-10$ & $--$ & $--$ \\ ${\cal O}_{qq}^{(1,3)}$ & $10$ & $-10$ & $--$ & $--$ \\ ${\cal O}_{qq}^{(8,3)}$ & $-24$ & $-10$ & $--$ & $--$ \\ \hline ${\cal O}_{dd}^{(1)}$ & $5.3$ & $-10$ & $400$ & $-16$ \\ ${\cal O}_{ud}^{(1)}$ & $-3.7$ & $-10$ & $400$ & $-16$ \\ ${\cal O}_{dd}^{(8)}$ & $5.3$ & $-10$ & $400$ & $-16$ \\ ${\cal O}_{ud}^{(8)}$ & $14$ & $-10$ & $400$ & $-16$ \\ \hline ${\cal O}_{qu}^{(1)}$ & $14$ & $-9.3$ & $400$ & $-15$ \\ ${\cal O}_{qu}^{(8)}$ & $-24$ & $-10$ & $400$ & $-16$ \\ ${\cal O}_{qd}^{(1)}$ & $9$ & $22$ & $--$ & $--$ \\ ${\cal O}_{qsd}^{(1)}$ & $5.4$ & $2.8$ & $-400$ & $-15$ \\ ${\cal O}_{qsd}^{(8)}$ & $5.3$ & $-10$ & $400$ & $-16$ \\ \hline ${\cal O}_{qsq}^{(1)}$ & $16$ & $-14$ & $400$ & $-15$ \\ ${\cal O}_{qsq}^{(8)}$ & $24$ & $-2.8$ & $-400$ & $15$ \\ ${\cal O}_{qqs}^{(1)}$ & $-74$ & $4.2$ & $400$ & $-15$ \\ ${\cal O}_{qqs}^{(8)}$ & $14$ & $-9.3$ & $400$ & $-15$ \\ ${\cal O}_{qq}^{(1s)}$ & $29$ & $22$ & $--$ & $--$ \\ ${\cal O}_{qq}^{(8s)}$ & $29$ & $22$ & $--$ & $--$ \\ \hline \end{tabular} \label{t: result} \end{table} The bounds on the $p$-wave phases arise from the contributions of the operator to $\epsilon$, and are weaker than the bounds on the $s$-wave phases that arise from the contributions to $\epsilon^\prime$. For the operator basis that we have been using, the bounds on the different components are not independent. This, however, is not an important point because there is nothing special about this operator basis. We prefer to illustrate separately the bounds on each parity and isospin amplitude because it is possible to construct operators with definite parity and isospin. \section{Two-quark Operators of Dimension six} In addition to the four-quark operators of dimension six considered in the previous section, there are also two-quark operators of dimension six that can contribute to the processes under consideration \cite{buchmuller}. These operators are the $SU(3)\times SU(2) \times U(1)$ invariant versions of ``penguin'' operators that naively appear to be dimension five \cite{rujula}. There are two types of operators that contribute to ${\cal CP}$ violation in $|\Delta S|=1$ processes. The first one, in the notation of Ref.~\cite{buchmuller} is: \begin{equation} {\cal O}_{dG} = (\overline{q}\sigma_{\mu\nu}\lambda^a d)\phi G^a_{\mu\nu} \end{equation} The operator of interest to us is obtained when the scalar doublet, $\phi$, takes its vacuum expectation value. This leads to the effective Lagrangian (with the same overall normalization that we used before and $v \approx 246$~GeV): \begin{eqnarray} {\cal L}_p&=&{g^2 \over \Lambda^2}{v \over \sqrt{2}}\biggl[ \lambda_{ds}\overline{d}\sigma_{\mu\nu}\lambda^a\biggl({1+\gamma_5 \over 2}\biggr) sG^a_{\mu\nu} + \lambda^\star_{sd}\overline{d}\sigma_{\mu\nu}\lambda^a\biggl( {1-\gamma_5\over 2}\biggr)sG^a_{\mu\nu}\biggr]+{\rm h.~c.} \nonumber \\ &\equiv & {g^2 \over \Lambda^2}{v \over \sqrt{2}} \overline{d}\sigma_{\mu\nu}t^a(f_{pc}+\gamma_5 f_{pv})sG^a_{\mu\nu} +{\rm h.~c.} \label{penguinop} \end{eqnarray} There are also analogous operators where the Gluon field strength tensor is replaced by field strength tensors for electroweak gauge bosons. The matrix elements of these operators are suppressed by a power of $\alpha = 1/137$ with respect to the gluon operator and we will, therefore, neglect them. \subsection{Constraint on the parity conserving coupling} The parity conserving coupling $f_{pc}$ is constrained by the contribution of Eq.~\ref{penguinop} to the parameter $\epsilon$. Unlike the four-quark operators of the previous section, we cannot use vacuum saturation to compute the matrix elements of this operator. However, this is the same operator that arises in the Weinberg model of ${\cal CP}$ violation, and the analysis has been carried out by Donoghue and Holstein \cite{dohowe} using MIT bag model matrix elements. We can simply take over their results to find: \begin{equation} |\epsilon|_p \approx 1.5 \times 10^{5} |\xi| \biggl({M_W\over \Lambda}\biggr)^2|{\rm Im}f_{pc}| \label{epconp} \end{equation} This contribution to $\epsilon$ is due to long distance effects as those discussed in the previous section. In complete analogy we have introduced the parameter $\xi$ which takes the values $\xi = 0.12$ for $\rho =0.8$ and $\xi = -0.48$ for $\rho =1.2$. We find that the sensitivity of the result to the $SU(3)$ breaking parameters of the pole model is larger in this case than it was for the four-quark operators. \subsection{Constraint on the parity violating coupling} The constraint on the parity violating coupling $f_{pv}$ comes from an analysis of $\epsilon^\prime$. Just as we did for $f_{pc}$, we simply take over the results of Ref.~\cite{dohowe} with a suitable identification of the coupling. We find: \begin{equation} \biggl|{\epsilon^\prime \over \epsilon}\biggr|_p \approx 2.2 \times 10^5 \biggl({M_W \over \Lambda}\biggr)^2 |{\rm Im}f_{pv}| \label{eppconp} \end{equation} \subsection{Contribution to $A(\Lambda^0_-)$} Once again we use the fact that up to coupling constants, this operator is the same one appearing in the Weinberg model of ${\cal CP}$ violation. Its matrix elements using the MIT bag model can thus be taken from Ref.~\cite{donpa}. We find: \begin{eqnarray} \phi^s_1 &\approx & 7 \times 10^4 \biggl({M_W \over \Lambda} \biggr)^2 {\rm Im}f_{pv} \nonumber \\ \phi^p_1 & \approx & -8 \times 10^4 \biggl({M_W \over \Lambda} \biggr)^2 {\rm Im}f_{pc} \end{eqnarray} {}From these it follows that: \begin{equation} A(\Lambda^0_-)_p\approx -10^4 \biggl({M_W \over \Lambda} \biggr)^2 \biggl({\rm Im}f_{pc}+0.9{\rm Im}f_{pv}\biggr) \end{equation} In the Weinberg model this operator appears with $f_{pv}=-f_{pc}$ and there is a large cancellation between the two phases leading to a smaller value for $A(\Lambda^0_-)$ than would have been obtained from each phase individually \cite{donpa}. In our general operator analysis, the bounds from Eq.~\ref{epconp} and Eq.~\ref{eppconp} can be combined to obtain (with $\xi =-0.5$): \begin{eqnarray} \biggl({M_W \over \Lambda} \biggr)^2 {\rm Im}f_{pc} &<& 2.7 \times 10^{-8} \nonumber \\ \biggl({M_W \over \Lambda} \biggr)^2 {\rm Im}f_{pv} &<& 6.8 \times 10^{-9} \end{eqnarray} or in terms of the hyperon decay observable: \begin{equation} A(\Lambda^0_-) \leq \cases{ 3 \times 10^{-4} & Parity conserving operator \cr 6 \times 10^{-5} & Parity violating operator \cr} \label{respen} \end{equation} Before ending this section we should comment on one class of two-quark operators that we have not discussed. In the notation of Ref.~\cite{buchmuller} it is: \begin{equation} {\cal O}_{qG} = i(\overline{q}\lambda^a \gamma_\mu D_\nu d)\phi G^a_{\mu\nu} \label{other} \end{equation} and related operators with field strength tensors for electroweak gauge bosons instead of the gluon. These latter ones will have matrix elements suppressed by $\alpha$ compared to Eq.~\ref{other}. We have not found a simple way to estimate the matrix elements of this operators and for this reason we do not discuss them in detail. We do not expect the behavior of this type of operator to be significantly different from the others that we have discussed. \section{Summary and Conclusions} The minimal standard model of electroweak interactions is in extraordinary agreement with all experiments conducted so far, and there is no evidence for any new particles below $100$~GeV or so. In view of this, it is reasonable to assume that any new physics beyond the minimal standard model is associated with a scale $\Lambda \geq M_W$, and it is, therefore, possible to represent the low energy effects of any such new physics with an effective Lagrangian that respects the symmetries of the standard model. In this paper we have studied all the $|\Delta S| =1$, ${\cal CP}$ violating, operators that occur at dimension six. We have investigated the constraints that exist on the couplings of these operators from the measurements of $\epsilon$ and $\epsilon^\prime$, and estimated what their largest contribution to ${\cal CP}$ violation in $A(\Lambda^0_-)$ could be. The operators that we have discussed also contribute to CP conserving and flavor changing amplitudes. We might thus worry, that the constraints on the real part of the couplings are such, that it is not natural for the imaginary (${\cal CP}$ violating) part of the couplings to attain the upper bounds allowed by the values of $\epsilon$ and $\epsilon^\prime$. We briefly address this issue in this section. Consider the contributions to $K^0-\overline{K}^0$ mixing. If we fix Im~$\lambda_i$ to its maximum allowed value, we find that the constraint $2{\rm Re}M_{12,i} \leq \Delta m_K$ is also satisfied if: \begin{equation} {\rm Re}\lambda_i \leq {{\rm Im}\lambda_i \over 2 \sqrt{2}\epsilon} \end{equation} Therefore, the ${\cal CP}$ conserving constraint is also satisfied if both real and imaginary parts of the couplings are of the same size or if the imaginary part is smaller than the real part by a factor of $\epsilon$. The strongest constraints on flavor changing operators in the ${\cal CP}$ conserving case are known to come from $K^0-\overline{K}^0$ mixing. If we set the couplings to be of order one, we obtain a lower bound on the scale of new physics $\Lambda$ requiring that $2{\rm Re}M_{12,i} \leq \Delta m_K$. It is easy to check that with couplings and scales satisfying this bound, the new operators do not make any significant contributions to the real part of the amplitudes in $K \rightarrow \pi \pi$ or $\Lambda \rightarrow p \pi$. Therefore, we conclude that fixing the imaginary part of the couplings to their maximum allowed value is not in conflict with ${\cal CP}$ conserving constraints. In the minimal standard model we have estimated previously \cite{steger} that $A(\Lambda^0_-)$ is of the order of a few times $10^{-5}$. For the new physics considered in this paper we find that most of the operators would naturally induce contributions to $A(\Lambda^0_-)$ at the $10^{-5}$ level, making them indistinguishable from the minimal standard model (as long as precise calculations of the matrix elements are not available), and inaccessible to the search to be conducted by E871. However, we have also found that for certain operators, ${\cal O}_{qqs}^{(1)}$ and ${\cal O}_{dG}$, $A(\Lambda^0_-)$ could be as large as a few times $10^{-4}$. Given our crude estimate of the hadronic matrix elements involved, all our numerical results should be viewed with caution. Nevertheless, our results suggest that the search for ${\cal CP}$ violation in $A(\Lambda^0_-)$ at the $10^{-4}$ level of sensitivity that is expected for E871 is potentially very interesting. Our results also suggest that this measurement is complementary to the measurement of $\epsilon^\prime/ \epsilon$, in that it probes potential sources of ${\cal CP}$ violation at a level that has not been probed by the kaon experiments. This is particularly true for parity conserving interactions that do not contribute to $\epsilon^\prime$ and are only constrained by $\epsilon$. We conclude that it is possible for E871 to observe a ${\cal CP}$ violating signal at the $10^{-4}$ level. Our study indicates that if such a signal is observed, it would probably be evidence for physics beyond the minimal standard model. However, a reliable determination of hadronic matrix elements is necessary to reach any definite conclusion. \section*{Acknowledgements} The work of G.V. was supported in part by a DOE OJI award under contract number DE-FG02-92ER40730. The work of X.G.He was supported in part by DOE contract number DE-FG06-85ER40224. G.V. thanks the Institute of Theoretical Science at the University of Oregon for their hospitality while part of this work was performed. We are grateful to J. F. Donoghue and N. Deshpande for helpful discussions. \clearpage
\section{#1} \setcounter{equation}{0}} \newtheorem{theo}{Theorem}[section] \begin{document} \newtheorem{lem}[theo]{Lemma} \newtheorem{prop}[theo]{Proposition} \newtheorem{coro}[theo]{Corollary} \title{Controlled Geometry via Smoothing \footnote{ 1991 {\em Mathematics Subject Classification}. Primary 53C20.}} \author{Peter Petersen\thanks{Partially supported by NSF and NYI grants} \\ \and Guofang Wei\thanks {Partially supported by NSF Grant \# DMS9409166. }\\ \and Rugang Ye\thanks{Partially supported by NSF Grant \# DMS9401106} } \date{} \maketitle \begin{abstract} We prove that Riemannian metrics with a uniform weak norm can be smoothed to having arbitrarily high regularity. This generalizes all previous smoothing results. As a consequence we obtain a generalization of Gromov's almost flat manifold theorem. A uniform Betti number estimate is also obtained. \end{abstract} \newcommand{\mbox{Tr}}{\mbox{Tr}} \newcommand{\mbox{inj}}{\mbox{inj}} \newcommand{\mbox{vol}}{\mbox{vol}} \newcommand{\mbox{diam}}{\mbox{diam}} \newcommand{\mbox{Ric}}{\mbox{Ric}} \newcommand{\mbox{conj}}{\mbox{conj}} \newcommand{\mbox{Hess}}{\mbox{Hess}} \newcommand{\mbox{div}}{\mbox{div}} \sect{Introduction} An ultimate goal in geometry is to achieve a classification scheme, using natural geometric quantities to characterize the topological type or diffeomorphism type of Riemannian manifolds. While this grand scheme seems to be an impossible dream, its basic philosophy has been a driving force in many important developments in Riemannian geometry. The sphere theorems and various topological finiteness theorems are typical examples. These results are concerned with control of global topology of manifolds, and a crucial point therein is to control, uniformly, the local topology. Control of local topology often follows from control of local geometry. Here, by local geometry, we mean the local behavior of the metric tensor. On the other hand, control of local geometry is frequently also the essential ingredient for control of global geometry, such as in Cheeger-Gromov's compactness theorem and its various extensions, which can be named geometric finiteness theorems. Notice that some rudimental topological finiteness results are direct corollaries of geometric finiteness theorems. But the significance of the latter goes beyond this. In any case, control of local geometry is obviously a key topic. An interesting and important aspect of this topic is various degrees of control of local geometry needed or available in different situations. In \cite{p}, the first author introduced a sequence of norms which provide a certain quantitative measure for local geometric control. These norms can be defined either in terms of the $C^{k,\alpha}$-norms or the $L^{k,p}$-norms for functions, and they are defined on a given scale. For example, the $C^{k,\alpha}$-norm of a Riemannian manifold on scale $r$ is bounded, if it is covered by coordinate charts of size comparable to $r$ such that the metric tensor expressed in the coordinates is uniformly bounded in $C^{k, \alpha}$-norm, and that the coordinate transition functions are uniformly bounded in $C^{k+1, \alpha}$-norm. Note that the local topology is uniformly trivial if one of these norms on some scale is bounded. To admit richer topological and geometric structures under norm bounds, we shall introduce a weak version of these norms. The essential new feature is that we allow coordinate maps to have double points. In spirit, this is similar to replacing an injectivity radius bound by a conjugate radius bound. (Of course, e.g. a weak (harmonic) $C^{0, \alpha}$ bound is so weak that it is far from implying a conjugate radius bound.) Indeed, our basic theme is to try to find the minimal degree of control of local geometry under which interesting geometric and topological consequences can be drawn. Traditional geometric conditions such as curvature bounds imply various degrees of local geometric control, so we can think from their perspectives. Historically, sectional curvature bounds were the first to be systematically studied. They can roughly be compared with weak $C^2$-norm bounds, at least the latter imply the former. Since understanding of sectional curvature bounds has been reached on a good level, it is natural to try to find the minimal degree of local geometric control under which a metric can be approximated by metrics with sectional curvature bounds or weak $C^2$-norm bounds (or better bounds). In this paper, we present a result towards this goal, along with some applications. The local geometric control we need is as weak as a bound on the weak harmonic $C^{0, \alpha}$-norm, or a bound on the weak $L^{1, p}$-norm. These do appear to be the sought-after minimal degree of local geometric control in our set-up. To formulate the result precisely, we introduce the following classes of Riemannian manifolds. (The definition of the weak norms are given in \S 2.) \newline \noindent {\bf Definition}. Given $n \geq 2$, $0 < \alpha < 1$, $p > n$ and function $Q: (0, \infty) \rightarrow [0, \infty)$ which is nondecreasing in $r$ and satisfies $\lim_{r \rightarrow 0} Q(r) = 0$, we define \[ {\cal M}(n, \alpha, Q) = \left\{ (M, g) \left| \begin{array}{l} (M,g) \mbox{ is a complete Riemannian manifold}, \dim M = n, \\ \mbox{the weak harmonic}\ C^{0, \alpha}\ \mbox{norm}\ \|(M,g)\|^{W,h}_{C^{0,\alpha}, r} \leq Q(r)\\ \mbox{ for all positive }\ r \leq 1 \end{array} \right. \right\}, \] and \[ {\cal M}(n, p, Q) = \left\{ (M, g) \left| \begin{array}{l} (M,g) \mbox{ is a complete Riemannian manifold}, \dim M = n, \\ \mbox{the weak}\ L^{1,p}\ \mbox{norm}\ \|(M,g)\|^{W}_{L^{1,p}, r} \leq Q(r)\\ \mbox{ for all positive}\ r\leq 1 \end{array} \right. \right\}. \] \noindent {\em Remark} 1. By Anderson-Cheeger's work \cite{ac} manifolds with a lower bound for Ricci curvature and a positive lower bound for conjugate radius belong to these two classes. \noindent {\em Remark} 2. It is unknown whether (weak) $L^{1,p}$ bounds imply (weak) harmonic $C^{0, \alpha}$ bounds. (By the Sobole v embedding, they do imply $C^{0,\alpha}$ bounds, but not yet harmonic $C^{0, \alpha}$ bounds. ) In other words, it is unknown whether controlled $C^{0,\alpha}$ harmonic coordinates exist under the assumption of a (weak) $L^{ 1, p}$ bound. This question seems rather subtle. We plan to return to it in the future. At the moment, we consider the said two bounds as independent conditions. Note that one can further ask whether (weak) $C^{0, \alpha}$ bounds imply (weak) harmonic $C^{0, \alpha}$ bounds. There does not seem to be any evidence to support an answer in the affirmative. \\ \begin{theo} \label{main} For every manifold $(M,g)$ in ${\cal M}(n, \alpha, Q)$ and positive numbers $\epsilon $, there are metrics $g_\epsilon$ on $M$ such that \begin{eqnarray} e^{-\epsilon}g \leq & g_\epsilon &\leq e^\epsilon g \label{t1},\\ \| (M,g_\epsilon)\|^{W}_{C^{0,\alpha}, r} & \leq & 2Q(r) \label{t2}, \\ \| (M,g_\epsilon)\|^W_{C^{k,\alpha}, r} & \leq & \widetilde{Q}, \end{eqnarray} \noindent where $k$ is an arbitary positive integer, $0 < r \leq 1$ and $\widetilde{Q} = \widetilde{Q} (n, k , \epsilon, \alpha, Q(r))$ denotes a positive number depending only on $n, k, \epsilon, \alpha$ and $Q(r)$. \end{theo} \begin{theo} \label{main'} For every closed manifold $(M,g)$ in ${\cal M}(n, p, Q)$ and positive numbers $\epsilon $, there are metrics $g_\epsilon$ on $M$ such that \begin{eqnarray} e^{-\epsilon}g \leq & g_\epsilon &\leq e^\epsilon g \label{t1'},\\ \| (M,g_\epsilon)\|^{W}_{L^{1,p}, r} & \leq & 2Q(r) \label{t2'}, \\ \| (M,g_\epsilon)\|^W_{L^{k,p}, r} & \leq & \widetilde{Q}, \end{eqnarray} \noindent where $k$ is an arbitary positive integer, $0 < r \leq 1$ and $\widetilde{Q} = \widetilde{Q} (n, k , \epsilon, p, Q(r))$ denotes a positive number depending only on $n, k, \epsilon, p$ and $Q(r)$. \end{theo} \noindent {\em Remark} Theorem 1.2 actually also holds for complete, non-compact manifolds. This will be shown in a paper by the third author. \\ Thus a metric with some regularity (given by the weak norm) can be deformed or smoothed to a nearby one with arbitrarily high regularity. In particular, {\it manifolds with a lower bound on Ricci curvature and a positive lower bound on conjugate radius can be smoothed.} Previous smoothing results have been concerned with metrics with various curvature bounds, and involved two independent techniques: the embedding method and the Ricci flow. The embedding technique in smoothing as used by Cheeger-Gromov \cite{cg} consists of embedding (or immersing) a given manifold into a Euclidean space and then perturbing it suitably by a smoothing operation, which is based on the classical convolution process. The smoothing result in \cite{cg} is that metrics on closed manifolds with lower and upper bounds on sectional curvatures and a positive lower bound on injectivity radius can be smoothed to metrics with bounds on all derivatives of the Riemann curvature tensor. Later, by embedding into a Hilbert space instead of a finite dimensional space, Abresch \cite{a} was able to remove the condition on injectivity radius and extend to complete manifolds. More recently Shen \cite{sh} showed that manifolds with a lower bound on sectionl curvatures and a positive lower bound on injectivity radius can be smoothed to having two-sided sectional curvature bounds. The technique of Ricci flow is based on the fundamental work of Hamilton \cite{ha}. Using this technique, Bemelmans-Min Oo-Ruh \cite{bmr} obtained the same result as in \cite{cg} without injectivity radius lower bound, and Shi \cite{s} obtained the same result as in \cite{a}. Later work considers metrics with other kinds ! of curvature bound. For example, in \cite{y1,y2} Yang dealt with integral bounds on sectional curvatures. In \cite{dwy}, Ricci curvature bounds were treated. By virtue of the available constructions of controlled harmonic coordinates under various curvature bounds, all these smoothing results are consequences of Theorem~\ref{main} or Theorem~\ref{main'}. As typical applications we present the following two results. \begin{theo}[Betti number estimate] For the class of manifolds $M^n$ in \\ ${\cal M}(n, \alpha, Q)$ or in ${\cal M}(n, p, Q)$, and satisfying $\mbox{diam}_M \leq D$, we have the estimate for the Betti numbers \begin{equation} \sum_i b^i(M^n) \leq C(n,D,\alpha,Q)\ \mbox{or}\ C(n,D,p,Q),\label{be} \end{equation} and the estimate for the number of isomorphism classes of rational homotopy groups \begin{equation} \pi_q(M) \otimes Q \leq C(n,q,D,\alpha,Q)\ \mbox{or}\ C(n,q,D,p,Q)\ \mbox{for}\ q \geq 2. \label{hge} \end{equation} \end{theo} (\ref{be}) follows from Theorem~\ref{main}, \ref{main'} and Gromov's uniform betti number estimate regarding sectional curvature \cite{g2}. This estimate can also be proved directly using Toponogov type comparison estimate introduced in \cite{w}, see \cite{pw} for details. In \cite{w} the same estimate (\ref{be}) is given for the class of manifolds satisfying $\mbox{Ric}_M \geq -(n-1)H$, conj $\geq r_0$ and $\mbox{diam}_M \leq D$. (\ref{hge}) follows from Theorem~\ref{main}, \ref{main'} and the results in \cite{r}. \begin{theo} \label{aflat} There exists an $\epsilon =\epsilon(n,\alpha,Q)$ or $\epsilon(n,p,Q)> 0$ such that if a manifold $M^n$ belongs to ${\cal M}(n, \alpha, Q)$ or ${\cal M}(n, p, Q)$ and diam $\leq \epsilon$, then $M$ is diffeomorphic to an infranilmanifold. \end{theo} This generalizes Gromov's almost flat manifold theorem \cite{g1} as well as its generalization in \cite{dwy}. (The proof is simple: combine Theorem 1.1 with \cite{g1}.) The proof of Theorem~\ref{main} and Theorem 1.2 uses the embedding method in \cite{a}. Roughly speaking, we embed a given Riemannian manifold into the Hilbert space of $L^2$-functions on it, and then use the embedding to pull back the $L^2$-metric of the Hilbert space. The crucial point is of course to find a suitable embedding, such that the pull-back metric will enjoy nice properties. In \cite{a}, the embedding is defined in terms of distance functions. In our situation, these functions are not appropriate, and we employ instead solutions of a canonical geometric partial differential equation. Now if e.g. the harmonic $C^{0,\alpha}$-norm of the manifold is bounded, then a uniform pointwise bound on sectional curvatures will hold for the pull-back metric, and hence we can apply the smoothing results for metrics with sectional curvature bounds as given e.g. in \cite{a} or \cite{s}. If we only assume that the weak harmonic $C^{0, \alpha}$-norm of the manifold is bounded, i.e. it is in the class ${\cal M}(n, \alpha, Q)$, the global embedding is generally not under control. To remedy the situation, we follow the idea in \cite{a} of employing instead local embeddings. In \cite{a}, Abresch uses the exponential map to lift local patches of the manifold and his local embeddings are exactly embeddings of these lifted patches. In our situation, the exponential map is not suitable. Our substitute for it is the coordinate maps. Thus we use them to lift local patches, and construct embeddings of the lifted patches via the same geometric partial differential equation as mentioned before. To make sure that the pull-back metrics induced by these local embeddings descend to the local patches and that the resulting metrics patch together to define a metric globally, it is crucial to require the embeddings to be equivariant under isometries. Since our embeddings are defined in terms of solutions of a canonical geometric PDE, they naturally share this equivariance property. Basically, the above scheme also works for manifolds in the class ${\cal M}(n, p, Q)$, but some modifications are necessary. As before, the said pull-back metrics descend to yield a new metric on the underlying manifold. But these metrics satisfy here an integral bound on sectional curvatures rather than a poinwise bound. This is a new situation. To handle it, we apply the Ricci flow and follow the arguments in \cite{dwy}. A pointwise bound on Ricci curvature is used in several places in\cite{dwy}. Since no such bound is available in our current situation, the arguments in \cite{dwy} need to be improved and modified. The result we thus arrive at not only completes the smoothing scheme for the class ${\cal M}(n, p, Q)$, but also provides some new understanding of short time existence of the Ricci flow. \sect{Norm, Weak Norm and Smoothing} Fix an integer $k \geq 0$ a number $0 \leq \alpha \leq 1$. The {\it $C^{k,\alpha}$-norm } of an $n$-dimensional Riemannian manifold $(M,g)$ on scale $r$, $\|(M,g)\|_{C^{k,\alpha}, r}$, is defined to be the infimum of positive numbers $Q$ such that there exist embeddings: $$\varphi_s: B(0,r) \subset R^n \rightarrow U_s \subset M$$ ($B(0,r)$ denotes the closed ball of radius $r$ centered at the origin) with images $U_s$, $s \in \cal S$ (an index set), with the following properties:\\ 1) $e^{-Q} \delta_{ij} \leq g_{s,ij} \leq e^Q \delta_{ij}$, \\ 2) Every metric ball $B(p, \frac{r}{10}e^{-Q}),\ p\in M$ lies in some set $U_s$,\\ 3) $r^{|j|+ \alpha} \|\partial^j g_{s,ij}\|_{C^\alpha} \leq Q$ for all multi-indices $j$ with $0 \leq |j| \leq k$. \\ Here $g_{s,ij}$ denote the coefficients of $g_s=\varphi_s^*g$ on $B(0,r)$, and $\delta_{ij}$ are the Kronecker symbols. Note that this definition is slightly different from the corresponding one in \cite{p}, where in addition the (rescaled) $C^{k+1, \alpha}$-norm of the transition functions are required to be under control. For convenience, we can call the $C^{k, \alpha}$-norm (of Riemannian manifolds) as defined in \cite{p} the {\it strong $C^{k, \alpha}$-norm}. (Note however that the "strong" harmonic $C^{k, \alpha}$-norm is equivalent to the harmonic $C^{k, \alpha}$-norm.) We define the harmonic $C^{k,\alpha}$-norm on scale $r$, $\|(M,g)\|^h_{C^{k, \alpha}, r}$, by requiring additionally the following\\ 4) $\varphi_s^{-1}: U_s \rightarrow R^n$ is harmonic, \\ which is equivalent to saying that \\ 4$'$) id $: B(0,r) \rightarrow B(0,r)$ is harmonic with respect to $g_{s}$ on the domain and the Euclidean metric on the target, which is in turn equivalent to saying that $$\sum_i \partial_i (g_s^{ij} \sqrt{ \det g_{s,ij}})= 0$$ for all $j$. If $k \geq 1$ and $p >n$ (when $k=1$) or $p > \frac{n}{2}$ (when $k \geq 2$), then we define the $L^{k,p}$-norm on the scale of $r$, $\|(M,g)\|_{L^{k,p}, r}$, by retaining 1) and 2), and replacing 3) by\\ 3) $r^{|j|-\frac{n}{p}} \|\partial^j g_{s,ij}\|_{L^p} \leq Q$ for all $1 \leq |j| \leq k$. The harmonic $L^{k,p}$-norm is defined similarly. For any choice of these norms, it is clear that the local topology is trivial on some uniform scale for any class of manifolds with uniformly bounded norm. (Note that the injectivity radius may not be uniformly positive though.) To allow nontrivial local topology, we introduce the weak norms $\|\ \ \|^W_{C^{k,\alpha},r}$ and $\|\ \ \|^W_{L^{k,p},r}$, which are defined in identical ways except that each $\varphi_s: B(0,r) \rightarrow U_s$ is assumed to be a {\it local} diffeomorphism instead of diffeomorphism. The corresponding weak harmonic norms $\|\ \ \|^{W,h}_{C^{k,\alpha},r}$ and $\| \ \ \|^{W,h}_{L^{k,p},r}$ are defined in a similar way, with 4) being replaced by 4$'$). Note that (weak) harmonic norms dominate (weak) norms on the same scale. We also have $\|\ \ \|^W_{\ ,r} \leq \|\ \ \|_{\ ,r}$ and $\|\ \ \|^{W,h}_{\ ,r} \leq \|\ \ \|^h_{\, r}$. All norms are continuous and non-decreasing in $r$. If $(M,g)$ is sufficiently smooth, these norms converge to zero as $r \rightarrow 0$. Furthermore, (weak) $C^{k,\alpha}\ (L^{k,p})$ norms vary continuously in the $C^{k,\alpha}\ (L^{k,p})$ topology of Riemannian manifolds. See \cite{p} for the relevant details. We point out that $R^n$ is the only space with norm $=0$ on all scales. And flat manifolds are the only spaces with weak norm $=0$ on all scales. Conventional geometric conditions such as curvature bounds imply norm bounds. Such implications are mostly contained in constructions of controlled harmonic coordinates and are a crucial ingredient for various compactness theorems. To have a clear perspective, we collect these results in the following proposition. \begin{prop} There is a $Q(H,i_0,r,p)$ with $\lim_{r \rightarrow 0} Q(H,i_0,r,p) =0$ such that for manifolds with \\ a) $|K| \leq H, \ \mbox{inj} \geq i_0$, then $\|(M,g)\|^h_{L^{2,p},r} \leq Q(H,i_0,r,p)$; \\ b) $|K| \leq H$, then $\|(M,g)\|^{W,h}_{L^{2,p},r} \leq Q(H,r,p)$; \\ c) $|\mbox{Ric} | \leq (n-1)H,\ \mbox{inj} \geq i_0$, then $\|(M,g)\|^h_{L^{2,p},r} \leq Q(H,i_0,r,p)$; \\ d) $\mbox{Ric} \geq -(n-1)H,\ \mbox{inj} \geq i_0$, then $\|(M,g)\|^h_{L^{1,p},r} \leq Q(H,i_0,r,p)$; \\ e) $\mbox{Ric} \geq -(n-1)H$, conj $ \geq i_0$, then $\|(M,g)\|^{W,h}_{L^{1,p},r} \leq Q(H,i_0,r,p)$. \end{prop} These results follow from works of Jost-Karcher \cite{jk}, Anderson \cite{an} and Anderson-Cheeger \cite{ac}. We now turn to the smoothing question. As explained in the introduction, our strategy is to first achieve sectional curvature bounds by embedding into the Hilbert space of $L^2$-functions. This is done in the next two sections. The higher regularity smoothing then easily follows from known smoothing results. Consider $(M,g) \in {\cal M}(n, \alpha, Q)$. We have a collection of local diffeomorphisms \[ \varphi_s:\ B(0,r) \rightarrow U_s \subset M \] satisfying 1), 2), 3) and 4$'$). In the next section we will construct a canonical embedding \[ F_s: (B(0,r),g_s) \rightarrow L^2 (B(0,r), g_s), \] where $g_s = \varphi_s^*g$. We use $F_s$ to pullback the $L^2$ metric of $L^2 (B(0,r), g_s)$ to produce a new metric $\tilde{g}_s$ on $B(0,r)$. This construction works for general metrics on $B(0,r)$, and has the following equivariance property, which will be proved in \S 4. Namely, if $g_1,\ g_2$ are two metrics on $B(0,r)$ such that there is an isometric embedding \[ \psi: (B(0,r), g_1) \rightarrow (B(0,r),g_2) \] and if $\tilde{g}_1,\ \tilde{g}_2$ are obtained via the above construction, then \[ \psi: (B(0,r), \tilde{g}_1) \rightarrow (B(0,r), \tilde{g}_2) \] is also an isometric embedding. Granted this (see Proposition~\ref{4.4}) we have \begin{prop} There exists a smooth metric $\bar{g}$ on $M$ such that the pullback of $\bar{g}$ by $\varphi_s$ is exactly $\tilde{g}_s$. \end{prop} \noindent {\em Proof.} Let $r_1= \frac{r}{10} e^{-Q}$. Then for every $p \in M$, $B(p,r_1) \subset U_s$ for some $s$. It follows that there exists a $\tilde{p} \in B(0,r)$ such that $B_{g_s}(\tilde{p},r_1) \subset B(0,r)$ and $\varphi_s (\tilde{p}) = p$. We now define the metric $\bar{g}$ as follows. If $X,Y \in T_pM$, then \[ \bar{g}(X,Y) = \tilde{g}_s \left(((\varphi_s)_*|_{\tilde{p}})^{-1}(X), ((\varphi_s)_*|_{\tilde{p}})^{-1}(Y)\right). \] To show that this metric is well-defined, let $\tilde{p}'$ be another such point, i.e. for some $s'$, $B_{g_{s'}}(\tilde{p}',r_1) \subset B(0,r)$ and $\varphi_{s'} (\tilde{p}') =p$. Let $r_4 = \frac{r}{20} e^{-4Q}$ and $r_3 = \frac{r}{20} e^{-3Q}$. Denote $g_0$ the Euclidean metric. Then we can show that \begin{lem} \label{hom} There is an isometric embedding \[ \psi: (B_{g_0}(\tilde{p},r_4),g_s) \rightarrow (B_{g_{s'}} (\tilde{p}',r_3),g_{s'}). \] \end{lem} \noindent {\em Proof.} First $\psi$ can be defined as follows. Since $g_s$ is $e^Q$-quasi-isometric to $g_0$, \begin{equation} \label{r34} B_{g_0}(\tilde{p}, r_4) \subset B_{g_s} (\tilde{p}, r_3). \end{equation} For any point $q \in B(\tilde{p},r_4)$, connect $q$ to the center point $\tilde{p}$ with a curve $\tilde{\gamma}$ in $B_{g_0}(\tilde{p},r_4)$ such that the length of $\tilde{\gamma}$ $l_{g_s}(\tilde{\gamma}) < r_3$. Since \[ \varphi_s:\ (B_{g_s} (\tilde{p}, r_3), g_s) \rightarrow B(p,r_3) \] is a local isometry and $\varphi_s (\tilde{p}) = p$. From (\ref{r34}) $\varphi_s$ maps the curve $\tilde{\gamma}$ to a curve $\gamma$ in $B(p,r_3)$ starting with $p$ and $l(\gamma) < r_3$. Again since $\varphi_{s'}$ is a local isometry and $\varphi_{s'} (\tilde{p}') =p$. The curve $\gamma$ then can be lifted via $\varphi_{s'}$ to a curve in $B_{g_{s'}}(\tilde{p}',r_3)$ starting with $\tilde{p}'$. The other end point of this curve is defined to be the image of $q$. (Note that, in general, lifting can not be done for incompelete space. Here the map is a local isometry and the curve starts from the center, and we have control on the length of the curve and the size of the metric ball, so it will not hit the boundary during lifting.) Now we will show that $\psi$ is well-defined, i.e. the image is independent of the choices of the curve $\tilde{\gamma}$. If $\tilde{\gamma}_1$ is another curve in $B_{g_0}(\tilde{p},r_4)$ connecting $q$ to the center point $\tilde{p}$ with $l_{g_s! }(\tilde{\gamma}_1) < r_3$, we ca $g_s$ is $e^Q$-quasi-isometric to $g_0$. Then $\varphi_s$ maps $\tilde{H}(s,t)$ to a homotopy $H(s,t)$ in $B(p,2r_3)$ with $l(H(s,\cdot)) < 2r_3$ for each $s$. Therefore $H(s,t)$ can be lifted via $\varphi_{s'}$ to a homotopy in $B_{g_{s'}}(\tilde{p}',2r_3)$ starting with $\tilde{p}'$. By the (localized) homotopy lifting lemma the other end points are all the same. Therefore $\psi$ is well-defined. Next we show that $\psi$ is one-to-one. Let $r_2 = \frac{r}{20} e^{-2Q}$. Then $$B_{g_{s'}} (\tilde{p}',r_3) \subset B_{g_0}(\tilde{p}',r_2) \subset B_{g_{s'}} (\tilde{p}',\frac{1}{2}r_1).$$ Since $B_{g_0}(\tilde{p}',r_2)$ is an Euclidean ball one can construct ``inverse" $\phi$ similarly as above: \[ \phi: \ (B_{g_{s'}} (\tilde{p}',r_3),g_{s'}) \rightarrow (B_{g_s}(\tilde{p},\frac{1}{2}r_1),g_s). \] Thus $\psi$ is one-to-one. That $\psi$ is an isometric embedding follows from the construction. \hspace*{\fill}\rule{3mm}{3mm}\quad Now using the equivariance, we have \[ \psi^* \tilde{g}_{s'} = \tilde{g}_s. \] Therefore \begin{eqnarray*} \lefteqn{\tilde{g}_{s'}\left(((\varphi_{s'})_*|_{\tilde{p}'})^{-1}(X), ((\varphi_{s'})_*|_{\tilde{p}'})^{-1}(Y)\right)} \\ & = & \tilde{g}_{s'}\left((\psi)_* \circ ((\varphi_s)_*|_{\tilde{p}})^{-1}(X), (\psi)_* \circ ((\varphi_s)_*|_{\tilde{p}})^{-1}(Y)\right) \\ & = & \psi^* \tilde{g}_{s'} \left(((\varphi_s)_*|_{\tilde{p}})^{-1}(X), ((\varphi_s)_*|_{\tilde{p}})^{-1}(Y)\right) \\ & = & \tilde{g}_s \left(((\varphi_s)_*|_{\tilde{p}})^{-1}(X), ((\varphi_s)_*|_{\tilde{p}})^{-1}(Y)\right). \end{eqnarray*} To show that $\varphi_s^* \bar{g} = \tilde{g}_s$, consider \[ \varphi_s : \ B(0, \frac{9}{10}r) \rightarrow U_s. \] In particular, for any $\tilde{p} \in B(0, \frac{9}{10}r)$, $B_{g_s}(\tilde{p},r_1) \subset B(0,r)$, and therefore $\tilde{p}$ can be used to define the metric $\bar{g}$ at $\varphi_s (\tilde{p})$. It follows from the definition that \begin{equation} \label{mm} \varphi_s^* \bar{g} = \tilde{g}_s. \end{equation} Finally, note that the smoothness of the metric $\bar{g}$ is an immediate consequence of (\ref{mm}). \hspace*{\fill}\rule{3mm}{3mm}\quad \sect{Embedding I} We continue with the above manifold $(M, g)$. Let $\Omega = B(0, r) \subset R^n$ with a pull back metric $\varphi_s^*g$. For convenience, this metric will be denoted by $g$. It is easy to see that $\|(\Omega, g)\|^h_{C^{0, \alpha}, r} \leq Q(r)$. We are going to construct an equivariant embedding of $(\Omega, g)$ into $L^2(\Omega, g)$ by associating to every point $p \in \Omega$ a geometric function $f_p \in L^2(\Omega) \equiv L^2(\Omega, g)$, which depends nicely on $p$. A natural choice seems to be the distance function measured from $p$. Indeed, it is used by Abresch in \cite{a}. However, under our rather weak assumptions on the metric it is impossible to have uniform control of the second order derivative of the distance function, which is needed to ensure that the pull-back metric induced by the embedding satisfies a sectional curvature bound. In fact, one can not even expect differentiability of the distance function in balls of uniform size. Our substitute for the distance function is solutions of a canonical geometric partial differential equation. Those solution functions have the crucial equivariance ( like the distance functions) and enjoy better regularity. Many choices of ``canonical" PDE solutions are possible, e.g. in \cite{a} Green's function is suggested. But Green's function is inconvenient because of its singularity. We shall employ a very simple and nicely-behaved PDE. Denote \[ \Omega_1 = \Omega \setminus \cup_{q \in \partial \Omega} \overline{B_g(q, i_0)}, \] where $i_0 =\frac{r}{10}$. ($B^g(q, \cdot)$ denotes the closed geodesic ball of center $q$ and radius $\cdot$ measured in $g$.) Then for $s \in \Omega_1$, let $h_s \in L^{1,2}_0 (\Omega)$ be the unique weak solution of the following Dirichlet boundary value problem: \begin{equation} \label{h} \left\{ \begin{array}{rcll} \Delta h_s & = & -1 & \mbox{in} \ B_g(s,i_0) \\ h_s & \equiv & 0 & \mbox{on} \ \partial B_g(s,i_0). \end{array} \right. \end{equation} Here the Laplace operator is defined with respect to the metric $g$. The function $h_s$ will be extended to be zero outside the geodesic ball. Since the harmonic $C^{0,\alpha}$-norm of $(\Omega, g)$ is uniformly bounded, it is easy to see that a uniform Poincare inequality holds on the balls $B_g(s,i_0)$ with dependence on $i_0$. A simple integration argument then yields a uniform estimate of the Sobolev norm of $h_s$. Uniform interior $C^{2, \alpha}$ estimates then follow readily, because in harmonic coordinates the Laplace operator takes the form $\Delta = g^{ij} \partial_i \partial_j$. We also have a uniform $L^{\infty}$ estimate up to boundary, but it seems impossible to obtain better estimate up to boundary because the control of the geometry of the boundary is very weak. At a first glance this appears to threaten to destroy the embedding scheme. Fortunately we have a way to get around it. On the other hand, we can not obtain control of the dependence of $h_s$ on the center $s$. To remedy this, we shall take a suitable average of $h_s$ over $s$. The resulting new family of functions will depend nicely on the center. Now let us state a few basic properties of the functions $h_s$ in the following proposition, which will be proved at the end of this section. Here, as before, we work under the assumption $\|(\Omega, g)\|^h_{C^{0, \alpha}, r} \leq Q(r)$. \begin{prop} \label{bh} Let $\bar{h}_s(p)$ be the solution of equation (\ref{h}) with respect to the canonical Euclidean metric $g_0$ on the Euclidean ball $B_{g_0}(s,i_0)$. Then for any $\epsilon > 0$ and fixed $0<R<1$, there is an $r_0 = r_0 (\epsilon,R,Q) > 0$ such that if $i_0 \leq r_0$, \begin{eqnarray} |h_s(p) - \bar{h}_s(p)| & < & \epsilon i_0^2, \label{c1} \\ | \frac{\partial}{\partial p} h_s(p) - \frac{\partial}{\partial p} \bar{h}_s(p)| & < &\epsilon i_0 \label{c2} \end{eqnarray} for all $s$ and all $p$ with $d_{g_0}(s,p) \leq Ri_0$. It will follow from the proof that $B_{g_0}(s,Ri_0) \subset B_g(s, i_0)$ so that these estimate make sense. Also \begin{equation} \label{ub} | \frac{\partial^2}{\partial p^2} h_s(p) |\leq C(n, Q,R), \ \ |\frac{1}{i_0} \frac{\partial}{\partial p} h_s(p) |\leq C(n, Q,R). \end{equation} \end{prop} Note that \begin{equation} \bar{h}_s(p) = \frac{1}{2n} (i_0^2 - d_{g_0}^2(s,p)). \end{equation} Therefore $\frac{2n}{i_0^2} \bar{h}_s(p) \leq \frac{1}{5}$ when $d_{g_0}(s,p) \geq \sqrt{\frac{4}{5}}i_0$. Choosing $R = \frac{10}{11}$ in Proposition~\ref{bh}, we have $\frac{2n}{i_0^2} h_s(p) < \frac{1}{4}$ when $ \sqrt{\frac{4}{5}}i_0 \leq d_{g_0}(s,p) \leq \frac{10}{11}i_0$ and $i_0$ is sufficiently small. Let $\beta = \beta_n \in C^\infty_0 ([0,\infty))$ be the cut off function. \[ \beta_n(t) = \left\{ \begin{array}{ll} 0 & \mbox{if} \ 0 \leq t \leq \frac{1}{4} \\ B_n & \mbox{if} \ t \geq \frac{1}{2} \end{array} \right., \] where $B_n$ is a constant which will be determined later. Then $\beta \left( \frac{2n}{i_0^2}h_s(p)\right) = 0$ near the sphere $d_{g}(s,p) = \frac{9}{10}i_0$ for all $i_0$ small. (Note that $d_{g}$ converges to $d_{g_0}$ when $i_0 \rightarrow 0$.) We define a new function which is $\beta \left( \frac{2n}{i_0^2}h_s(p)\right)$ restricted to the ball $B(s, \frac{9}{10}i_0)$ and identically zero outside. For simplicity we still denote this new function by $\beta \left( \frac{2n}{i_0^2}h_s(p)\right)$. As mentioned before, we have no control of the dependence of $h_s$ on the center $s$. The said average function is given as follows \begin{equation} f_p(q) = \int_\Omega \beta \left( \frac{2n}{i_0^2}h_s(p)\right) \beta \left( \frac{2n}{i_0^2}h_s (q) \right)\, ds. \end{equation} Note that $f_p(q)$ is symmetric in $p$ and $q$ and is $C^{2,\alpha}$ uniformly bounded in both variables. Now we define the embedding \begin{eqnarray*} F: \ \Omega_1 & \rightarrow & L^2(\Omega, g)\\ p & \rightarrow & i_0^{-\frac{3}{2}n+1} f_p(q) \end{eqnarray*} Note that \begin{eqnarray} d_{v_p}F : & q \longmapsto & 2ni_0^{-\frac{3}{2}n}\int_\Omega \beta'\left( \frac{2n}{i_0^2}h_s(p)\right) \langle \frac{1}{i_0}\nabla_{v_p} h_s(p), v_p \rangle \beta \left( \frac{2n}{i_0^2}h_s (q) \right)\, ds, \label{dff}\\ \nabla^2_{v_p,w_p}F: & q \longmapsto & 4n^2i_0^{-\frac{3}{2}n-1}\int_\Omega \beta''\left( \frac{2n}{i_0^2}h_s(p)\right) \langle \frac{1}{i_0}\nabla h_s(p), v_p \rangle \langle \frac{1}{i_0}\nabla h_s(p), w_p \rangle \beta \left( \frac{2n}{i_0^2}h_s (q) \right)\, ds \nonumber \\ & & + 2ni_0^{-\frac{3}{2}n-1}\int_\Omega \beta'\left( \frac{n}{i_0^2}\right) \nabla^2_{v_p,w_p} h_s(p) (v_p,w_p) \beta \left( \frac{n}{i_0^2}h_s (q) \right)\, ds. \label{dff2} \end{eqnarray} We first show that when $\Omega$ is an Euclidean domain, we can normalize $\beta$ so that $F$ is an isometric imbedding. In this case the imbedding function \[ \bar{f}_p(q) = \int_\Omega \beta \left(1-\frac{d_{g_0}^2(p,s)}{i^2_0}\right) \beta \left(1-\frac{d_{g_0}^2(q,s)}{i^2_0}\right)\, ds.\] By the symmetry of the integration domain, $B(p, \frac{\sqrt{3}i_0}{2}) \cap B(q, \frac{\sqrt{3}i_0}{2})$, and the integrand, $\bar{f}_p(q)$ depends only on $d_{g_0}(p,q)$ (and $\beta$). We can write $\bar{f}_p(q) = i^n_0 \tilde{f} (n, \frac{1}{i_0}d_{g_0}(p,q))$, $ d_{v_p}\bar{F}(q) = i_0^{-n/2}\tilde{f}' (n, \frac{1}{i_0}d_{g_0}(p,q)) \langle \nabla d_{g_0}(p,q), v_p \rangle$. Then \begin{eqnarray*} \| d_{v_p}\bar{F}\|^2_{L^2(\Omega)} & = & i_0^{-n}\int_{B(p, 2i_0)} \tilde{f}'^2 (n, \frac{1}{i_0}d_{g_0}(p,q)) \langle \nabla d_{g_0}(p,q), v_p \rangle^2 dq \\ & = & i_0^{-n}\int_0^{2i_0} r^{n-1} \int_{S^{n-1}} \tilde{f}'^2 (n, \frac{1}{i_0}r) \langle \xi, v \rangle^2 d\xi dr \\ & = & i_0^{-n}\frac{\mbox{vol} (S^{n-1})|v|^2}{n}\int_0^{2i_0} r^{n-1}\tilde{f}'^2 (n, \frac{1}{i_0}r) dr \\ & =& \frac{\mbox{vol} (S^{n-1})|v|^2}{n}\int_0^{2}r^{n-1}\tilde{f}'^2 (n,r) dr. \end{eqnarray*} Choose $B_n$ in the defintion of $\beta$ so that $\frac{\mbox{vol} (S^{n-1})}{n}\int_0^{2}r^{n-1}\tilde{f}'^2 (n,r) dr = 1$. Then we will have achieved the following. \begin{lem} $\bar{F}$ is an isometric embedding. \end{lem} With the above choice of $\beta$ we will show that $F$ is an almost isometric embedding when $\Omega$ is not necessarily an Euclidean domain and the second derivative of $F$ is also uniformly bounded. More precisely we have \begin{prop} \label{e-df} For any given $\epsilon_0 >0$, there exists an $r_0>0$ such that \begin{equation} \label{df1} (1+ \epsilon_0)^{-2} |v|^2 \leq \|d_{v_p}F\|^2_{L^2(\Omega)} \leq (1+ \epsilon_0)^{2} |v|^2, \end{equation} for all $v\in T_p\Omega_1$ and $0 < i_0 \leq r_0$. And \begin{equation} \label{df2} \|\nabla^2_{v_p,w_p}F\|^2 _{L^2(\Omega)} \leq C(n,\alpha,Q) i_0^{-2}|v|^2 \cdot |w|^2. \end{equation} \end{prop} \noindent {\em Proof.} By definition \[ \|d_{v_p}F\|^2_{L^2(\Omega)} = \int_{B(p,2i_0)} |d_{v_p}F(q)|^2 dq. \] Now the volume element of the metric $g$ is comparable with Euclidean one. Namely \begin{equation} \label{vv} e^{-Q(i_0)} \mbox{vol}_{R^n} \leq \mbox{vol}_g \leq e^{Q(i_0)} \mbox{vol}_{R^n}. \end{equation} Therefore it suffices to prove that $|d_{v_p}F(q)|$ is close to $|d_{v_p}\bar{F}(q)|$ when $i_0$ is small, which follows from (\ref{dff}) and Proposition~\ref{bh}. (\ref{df2}) also follows from (\ref{dff2}), (\ref{vv}) and Proposition~\ref{bh}. \hspace*{\fill}\rule{3mm}{3mm}\quad \\ {\em Proof of} Proposition~\ref{bh}. First we introduce some new functions. Let $\tilde{h}_{s, i_0}(p)$ be the solutions of (\ref{h}) on $B_{i_0^{-2}g}(s,1)$ with respect to the scaled metrics $i_0^{-2}g$, and $\bar{\tilde{h}}_s(p)$ the solution of (\ref{h}) with respect to the Euclidean metric on the Euclidean ball $B(s,1)$. Then \begin{equation} \label{hh} \tilde{h}_{s,i_0}(p)= i_0^{-2} h_s(p), \ \ \bar{\tilde{h}}_s(p) = i_0^{-2} \bar{h}_s(p). \end{equation} (Here the variables $s,p$ are in domain with different metrics for different functions.) Since \begin{equation} \label{normb} \|(B_{i_0^{-2}g}(s,1), i_0^{-2}g)\|_{C^{0,\alpha},1} = \|(B_g(s,i_0), g)\|_{C^{0,\alpha},i_0} \leq Q(i_0) \leq Q(1), \ \mbox{for}\ i_0 \leq 1, \end{equation} for the same reasons as mentioned before for $h_s$, we have the following estimates \begin{equation} \label{el} \| \tilde{h}_{s,i_0}\|_{L_0^{1,2}(B_g(s,1))} \leq C(n,Q(1)) \end{equation} and for any fixed $0<R<1$, \begin{equation} \label{ell} \| \tilde{h}_{s,i_0} \|_{C^{2,\alpha}(B_g(s,R))} \leq C(n, Q(1),R). \end{equation} On the other hand, the hypothesis $\|(B_{i_0^{-2}g}(s,1), i_0^{-2}g)\|_{C^{0,\alpha},1} \leq Q(i_0)$ implies that \[ e^{-Q(i_0)} d_{i_0^{-2}g_0}(p,s) \leq d_{i_0^{-2}g}(p,s) \leq e^{Q(i_0)} d_{i_0^{-2}g_0}(p,s). \] Therefore \[ B_{i_0^{-2}g_0} (s, e^{-Q(i_0)}R) \subset B_{i_0^{-2}g} (s, R) \subset B_{i_0^{-2}g_0}(s, e^{Q(i_0)}R). \] {}From these estimates and the uniqueness of the weak solution $\bar{\tilde h}_s$ it is easy to deduce the following: for each sequence of centers $s_k$ converging to some center $s_0$ and each sequence $i_0(k)$ converging to zero, the corresponding rescaled solutions $\tilde h_{s_k, i_0(k)}$ converge weakly to $\bar{\tilde h}_{s_0}$. Moreover, by the Arzela-Ascoli theorem, they also converge uniformly in $C^1$ on proper compact subsets of $B_{g_0}(s_0, 1)$. This convergence fact along with the smooth dependence of $\bar{\tilde h }_s$ on $s$ then imply that the $\tilde h_{s, i_0}$ converge uniformly with respect to $s$ in $C^1$ on proper compact subsets of $B_{g_0}(s, 1)$ as $i_0$ goes to zero. Consequently, for a fixed $R \in (0,1)$, given any $\epsilon >0$, there is an $r_0 >0$ such that for all $s$ and $p$ with $d_{g_0}(p,s) < R$, if $i_0 \leq r_0$, then \begin{equation} |\tilde{h}_s(p) -\bar{\tilde{h}}_s(p)| < \epsilon. \label{pc1} \end{equation} Similarly, \begin{equation} | \frac{\partial}{\partial p} \tilde{h}_s(p) -\frac{\partial}{\partial p} \bar{\tilde{h}}_s(p)| < \epsilon. \label{pc2} \end{equation} Hence for all $s$ and all $p$ with $d_{g_0}(s,p) \leq Ri_0$, \begin{eqnarray*} |h_s(p) - \bar{h}_s(p)| & < & \epsilon i_0^2, \\ | \frac{\partial}{\partial p} h_s(p) - \frac{\partial}{\partial p} \bar{h}_s(p)| & < &\epsilon i_0. \end{eqnarray*} (\ref{ub}) just follows from (\ref{ell}) and (\ref{hh}). \hspace*{\fill}\rule{3mm}{3mm}\quad \sect{Embedding II} In this section we study the geometry of $F(\Omega_1)$ as a submanifold in $L^2(\Omega)$. We will prove, among other things, two important properties of $F(\Omega_1)$. That is, the induced metric of $F(\Omega_1)$ has uniformly bounded sectional curvature and the embedding $F$ is equivariant. The geometry of $F(\Omega_1)$ is completely determined by the second fundamental form of its embedding into $L^2(\Omega)$, which in turn can be described by the family of orthogonal projections. $P(y):\ L^2(\Omega) \rightarrow T_yF(\Omega_1) \subset L^2(\Omega), y \in F(\Omega^1)$. We have \begin{lem} \label{4.1} The sectional curvature of $F(\Omega_1)$ is given by the following formula: \begin{equation} \label{curv} R(z_1, z_2) z_3 = [d_{z_1} P, d_{z_2}P] z_3, \ \ z_1,z_2,z_3 \in T_z F(\Omega_1). \end{equation} \end{lem} \noindent {\em Proof.} Since $P^2 =P$, one has \begin{equation} \label{p2} (d_{z_1}P)P + P (d_{z_1}P) = d_{z_1}P. \end{equation} Let $\nabla$ be the connection on $F(\Omega_1)$ and $d_{z_1}$ the directional derivative on the $L^2$ space. Then \begin{eqnarray*} \nabla_{z_1} z_2 & = & P(d_{z_1}z_2) \\ & = & d_{z_1}z_2 - (1- P)(d_{z_1}(Pz_2)) \\ & = & d_{z_1}z_2 - (1-P) \left[ (d_{z_1}P)z_2 + P (d_{z_1}z_2) \right] \\ & = & d_{z_1}z_2 - (1-P) (d_{z_1}P) (Pz_2) \\ & = & d_{z_1}z_2 - (d_{z_1}P)z_2. \end{eqnarray*} Here we have used (\ref{p2}) in the last equation. Therefore \begin{equation} \label{p1} \nabla_{z_1} = d_{z_1} - (d_{z_1}P). \end{equation} Now formula \ref{curv} follows from (\ref{p1}) and the definition of the curvature tensor. \hspace*{\fill}\rule{3mm}{3mm}\quad \begin{prop} \label{4.2} Let $\alpha_0 = (1+\epsilon_0)^2 C(n,\alpha,Q)$. Here $\epsilon_0, r_0, C$ are the same constants as in Proposition~\ref{e-df}. Then for all $0 < i_0 < r_0$, \begin{equation} \label{ep} \| d_{\dot{y}}P\|_{op} \leq \alpha_0 i_0^{-1} \| \dot{y} \|, \end{equation} \end{prop} \noindent {\em Proof.} Since $\left( 1-P(F(p)) \right) d_{w_p} F = 0$, \[ d_{d_{v_p}F} P \cdot d_{w_p} F = \left( 1-P(F(p)) \right) \nabla^2_{v_p,w_p}F. \] By (\ref{df1}) and (\ref{df2}), $\| d_{\dot{y}}P\|_{op} \leq \alpha_0 i_0^{-1} \| \dot{y} \|$. \hspace*{\fill}\rule{3mm}{3mm}\quad \\ Therefore the metric $\tilde{g} = F^* g_{L^2}$, the metric on $\Omega_1$ obtained by pulling back the $L^2$ metric, has bounded sectional curvatures. To prove the equivariance, we first note: \begin{lem} \label{uh} Let $h_s(p)$ be the function defined in (\ref{h}), and let $\psi: \Omega \rightarrow \Omega'$ be an isometric embedding. Then \begin{equation} h_{\psi (s)} (\psi (p)) = h_s(p). \end{equation} \end{lem} \noindent {\em Proof.} Since equation (\ref{h}) is invariant under isometry, this follows from the uniqueness of solutions to (\ref{h}). \hspace*{\fill}\rule{3mm}{3mm}\quad \\ Let $(\Omega,g)$ be as before and $F: \Omega_1 \rightarrow L^2(\Omega)$ the embedding defined in \S 3. With the above lemma, we can now prove \begin{prop} \label{4.4} If $\psi : (\Omega,g) \rightarrow (\Omega',g')$ is an isometric embedding, then $$\psi : (\Omega, \tilde{g}) \rightarrow (\Omega', \tilde{g}')$$ is also an isometric embedding. \end{prop} \noindent {\em Proof.} First, we assume $\psi$ is actually an isometry. Then \[ F \circ \psi (p) = f_{\psi (p)}, \] where the function \begin{eqnarray*} f_{\psi (p)} (q) & = & \int_\Omega \beta \left( \frac{2n}{i_0^2}h_s(\psi (p))\right) \beta \left( \frac{2n}{i_0^2}h_s (q) \right)\, ds. \\ & = & \int_\Omega \beta \left( \frac{2n}{i_0^2}h_{\psi^{-1}(s)}(p)\right) \beta \left( \frac{2n}{i_0^2}h_{\psi^{-1}(s)} (\psi^{-1}(q)) \right)\, ds. \end{eqnarray*} Here we have used Lemma~\ref{uh}. Since $\psi$ is an isometry, a change of coordinates yields \[ f_{\psi (p)}(q) = f_p (\psi^{-1} (q)). \] It follows then that \[ F \circ \psi = (\psi^{-1})^* \circ F, \] where we have denoted by $(\psi^{-1})^*$ the map on $L^2(\Omega)$ induced by $\psi^{-1}$. Therefore \[ \psi^* F^*g_{L^2} = (F \circ \psi)^* g_{L^2} = F^* ((\psi^{-1})^*)^* g_{L^2} = F^*g_{L^2}. \] This proves the equivariance when $\psi$ is an isometry. Since $\Omega, \Omega'$ are both domains of $R^n$, the general statement follows by applying the above to $\psi:\ \Omega \rightarrow \psi (\Omega)$. \hspace*{\fill}\rule{3mm}{3mm}\quad \\ {\it Proof of} Theorem 1.1. This theorem is a consequence of Proposition 2.2, Lemma 4.1, Proposition 4.2 and Proposition 3.3. \hspace*{\fill}\rule{3mm}{3mm}\quad \sect{Proof of Theorem 1.2} We consider $(M,g) \in {\cal M}(n,p,Q)$ and $\Omega = B(0, r) \subset R^n$ with the pull-back metric $\varphi_s^*g$, where $\varphi_s$ is a coordinate map. For convenience, we shall again denote the pull-back metric by $g$. We have the inequality $\|(\Omega, g)\|_{L^{1,p}, r} \leq Q(r)$. We employ the same embedding of $(\Omega, g)$ as before. Thus we use the same functions $h_s \in L_0^{1,2}(\Omega)$, as given by (3.1). But the estimates for $h_s$ are different now. By the $L^p$ elliptic theory we have uniform $L^{2,p}$ estimates for $h_s$ in the interior. A result similar to Proposition 3.1 then holds, namely we have the estimates (3.2), (3.3) and the second one in (3.4), while the first one in (3.4) is replaced by an estimate on the $L^p$-norm of the second order derivative. Now it is clear that our smoothing process produces a metric $\bar g$ on $M$ such that the lifted metrics $\varphi_s^* \bar g$ on $B(0, r/2)$ have uniformly bounded Sobolev constant and uniformly $L^p$-bounded sectional curvatures. Next we apply the Ricci flow to deform $\bar g$. For this purpose, we assume that $M$ is closed. We appeal to the arguments in \cite{dwy}. There, manifolds with a pointwise bound on Ricci curvature and a conjugate radius bound are treated. These conditions are used to show that controlled harmonic coordinates exist on lifted local patches, where the lifting is given by the exponential map. In these coordinates, the Ricci curvature bound then implies an $L^p$-bound on sectional curvatures. In our situation, we do have an $L^p$-bound on sectional curvatures. But the Ricci curvature bound is also used in several other places in \cite{dwy}. Since this bound is not available here, we need to modify the arguments in \cite{dwy}. In the key Proposition 3.1 (uniform short time existence of the Ricci flow with a priori control) in \cite{dwy}, we drop the estimate (3.4) on Ricci curvature. (Note that the conclusion of the proposition without (3.4) still suffices for our purpose.) We claim that the proposition then holds in our new situation. For convenience, we shall call this proposition the "Key Proposition". In \cite{dwy}, the proof of Key Proposition is based on four lemmata: Lemmata 3.2, 3.3, 3.4 and 3.5. Now let's take a look at these lemmata in our new situation. Lemma 3.2 (pointwise estimate for Riemann curvature tensor) holds without change. The proof of Lemma 3.3 ($L^p$-estimate for Riemann curvature tensor) depends on a {\it covering estimate}, i.e. an estimate for certain covering number and multiplicity regarding geodesic balls in the lifted patches. In \cite{dwy}, the estimate comes from the Bishop-Gromov covering argument, which depends on a pointwise lower bound for Ricci curvature. Now we do not have such a bound. But we still have a covering estimate, which follows from the properties of our coordinates and the basic control over the pull-back metrics as given by (the modified versions of) Propositions 3.1 and 3.2 (in the present paper). Another ingredient in the proof of Lemma 3.3 is an isometry correspondance between geodesic balls on different lifted patches. Now it is given by Lemma~\ref{hom}. (The proof of this correspondence given in \cite{dwy} does not work here.) Thus Lemma 3.3 also holds. Since we have dropped the estimate (3.4) about Ricci curvature in Key Proposition, Lemma 3.4 is no longer needed. Finally, note that the proof for Lemma 3.5 (estimate of the Sobolev constant) in \cite{dwy} goes by computing the change rate of the Sobolev constant along the flow. In \cite{dwy}, this rate is controlled by a uniform bound on Ricci curvature, which is not valid here. However, Lemma 3.2 contains an estimate for Ricci curvature at positive time $t$, namely it is dominated by a constant times $t^{-1/2}$. Since the function $t^{-1/2}$ is integrable at $0$, it is clear that the change rate of the Sobolev constant is still under control without a uniform Ricci curvature bound, and hence Lemma 3.5 carries over. (An alternative way of handling the Sobolev constant is to apply Yang's estimate for it in \cite{y2}, which uses only an $L^p$-bound on Ricci curvature and a positive lower bound on (local) volume. But that is more involved. ) We leave to the reader to formulate precisely the independent result implied by the above proof about short time existence of the Ricci flow.
\section{Introduction} Studies of the dynamics of a scalar field together with other forms of matter usually assume that this behaves as a perfect fluid. Few authors have taken into account dissipative effects \cite{Sin88} \cite{Sin90} \cite{Roy92} \cite{Roy93}, even though processes like particle creation might have been important in the early universe. In this paper we study the evolution of a universe filled with a massless minimally coupled scalar field and a viscous fluid. We find exact solutions of the Einstein equations in a Robertson-Walker metric and we analyse their asymptotic stability by means of the Lyapunov method \cite{Kra}. \section{The model} We investigate the evolution of a universe filed with a scalar field and a viscous fluid. The scalar field $\phi $ is free and minimally coupled to gravity, so that it obeys the equation $\Box \phi =0$. In the case of the homogeneous isotropic Robertson-Walker metric \begin{equation}\label{1} ds^2=dt^2-a^2(t)\left [{\frac{dr^2}{1-kr^2}}+r^2(d\theta ^2+\sin {}^2\theta d\phi ^2)\right] \end{equation} \noindent the scalar field equation becomes \begin{equation} \label{100} \ddot \phi +3H\dot \phi =0 \end{equation} Besides, only the bulk viscosity needs to be considered. Thus we replace in the Einstein equations the equilibrium pressure $p$ by an effective pressure \cite{Wei71} \begin{equation}\label{2} 3H^2=\frac {1}{2}\dot\phi^2+\rho -3{\frac k{a^2}}+\Lambda \end{equation} \begin{equation} \label{3} 2\dot H+ 3H^{2} = -\frac{1}{2}\dot\phi^2 - p -\sigma-\frac{k}{a^2}+\Lambda \end{equation} \noindent where $H=\dot a/a$, $^{\cdot }=d/dt$, $\rho $ is the energy density, $\sigma $ is the viscous pressure, $\Lambda$ is the cosmological constant and we use units $c=8\pi G=1$. As equation of state we take \begin{equation} \label{4} p = ( \gamma -1 ) \rho \end{equation} \noindent with a constant adiabatic index $0\le\gamma \le 2$, and we assume that $\sigma $ has the constitutive equation \begin{equation}\label{45} \sigma = - 3 \zeta H \end{equation} \noindent where $\zeta \ge 0$ is the bulk viscosity coefficient. Thus, we must solve equation (\ref{100}) together with the Einstein equations (\ref{2})(\ref{3}). Equation (\ref{100}) has the first integral \begin{equation} \label{5} \dot\phi=\frac{C}{a^3}\qquad \end{equation} \noindent where $C$ is an arbitrary integration constant, and we are interested in the case that $C\neq0$. Then, adding (\ref{2}) and (\ref{3}), and eliminating $\rho$, we arrive at \begin{equation} \label{6} 2\dot H+3\gamma H^2-3\zeta H=-\frac{2-\gamma}{2}\frac{C^2}{a^6}+ \left(2-3\gamma\right)\frac{k}{a^2}+\gamma\Lambda \end{equation} To find exact solutions of this equation we make the change of variable $a=v^\nu$, and we consider two cases: $\zeta$ constant and $\zeta \sim H$. We assume now that $\zeta$ is a constant. This is a good approximation when the thermodynamical variables do not change too much. This has been studied in several papers \cite{Tre} \cite{Hel73} \cite{Hel74} \cite{Sus} \cite{Roy83}. Choosing $\nu=2/(3\gamma)$, the equation (\ref{6}) becomes \begin{equation} \label{7} \ddot v+\frac{3}{2} \zeta\dot v+M v^m+N v^n-\frac{\gamma\Lambda}{2\nu}v=0 \end{equation} \noindent where $M=(3/8)\gamma\left(2-\gamma\right)C^2$, $m=1-4/\gamma$, $N=(3/4)\gamma(3\gamma-2)k$, $n=1-4/(3\gamma)$. It is linear for $\gamma=2$ and $k=0$, so that we can find its general solution. In this case, the evolution of the scale factor is the same as in a model without the scalar field and a bulk viscosity coefficient $\zeta/2$ \cite{Tre} \cite{Hel73}. Thus, we just quote the solutions without further analysis. \bigskip \noindent Let us consider first that $\Lambda=0$. We find \bigskip \begin{equation} \label{8} a(t)=\left[A \exp\left(\frac{3}{2}\zeta t\right)+B\right]^{1/3} \end{equation} \begin{equation} \label{9} \Delta\phi(t)=\frac{2C}{3\zeta B}\ln\left[\frac{\exp\left(\frac{3}{2}\zeta t\right)} {A\exp\left(\frac{3}{2}\zeta t\right)+B}\right] \end{equation} \begin{equation} \label{10} \rho(t)=\frac{\frac{3}{2}\zeta^2 A^2\exp(3\zeta t)-C^2} {2\left[A\exp\left(\frac{3}{2}\zeta t\right)+B\right]^2} \end{equation} \noindent where $A$ and $B$ are arbitrary integration constants. The requirement $\rho\ge0$ implies that there is a minimum time for the validity of this solution. Thus it describes a geodesically incomplete manifold, which in this sense is singular \cite{Ger}. Clearly, in its asymptotically de Sitter regime it is physically justified the approximation of constant $\zeta$. However, for shorter times, this assumption may not be justified. The asymptotically de Sitter stage occurs for any other value of $\gamma$ and $k$. In effect, assuming $a\sim\exp(H_0 t)$ for $t\to\infty$, with $H_0$ a constant, we find that $H_0=\zeta/\gamma$. To verify the asymptotic stability of this solution we use $a$ as the independent variable and we take the Liapunov function $L=(H-H_0)^2$. Then, we find that \begin{equation} \label{11} L'=-3\gamma\frac{L}{a}+O\left(\frac{1}{a^3}\right) \end{equation} \noindent is negative for large times. Let us consider now that $\Lambda\neq0$. There is a critical value for the cosmological constant $\Lambda_0=-3\zeta^2/16$. Thus, we must distinguish several cases: \noindent Two kinds of solutions appear for $\Lambda>\Lambda_0$. Singular evolutions \begin{equation} \label{12} a(t)=a_0 e^{3\Delta t/4}\left[\sinh\left(\omega \Delta t \right)\right]^{1/3} \end{equation} \noindent or nonsingular evolutions \begin{equation} \label{13} a(t)=a_0 e^{3\Delta t/4}\left[\cosh\left(\omega \Delta t \right)\right]^{1/3} \end{equation} \noindent where $\omega=\left[3\left(\Lambda-\Lambda_0\right)\right]^{1/2}$, $\Delta t=t-t_0$ and $a_0$, $t_0$ are integration constants. For the scalar field we find \begin{equation} \label{14} \Delta \phi(t)=-\frac{C}{B\mu}e^{(\omega-3\zeta/4)\Delta t} {}_2F_1\left(1,\frac{1}{2}-\frac{3\zeta}{8\omega},\frac{3}{2} -\frac{3\zeta}{8\omega},-\frac{A}{B}e^{2\omega \Delta t}\right) \end{equation} While $\rho\to (3/8)\zeta^2+(3/2)\zeta\omega>0$ for $t\to\infty$, in the nonsingular solutions $\rho$ becomes negative before a minimum time. On the other hand, for singular solutions, we find \begin{equation} \label{15} \rho\sim\left(\frac{1}{3}-\frac{C^2}{2a_0^3\omega^2}\right)\frac{1}{\Delta t^2} \qquad \Delta t\to 0 \end{equation} \noindent so that the energy density may remain positive along all the evolution. \noindent In the case $\Lambda=\Lambda_0$, the evolution is always singular \begin{equation} \label{16} a(t)=a_0 (\Delta t)^{1/3} e^{3\Delta t/4} \end{equation} \noindent In the case $\Lambda\le\Lambda_0$, the scale factor recollapses to a second singularity \begin{equation} \label{17} a(t)=a_0 e^{3\Delta t/4}\left[\sin\left(|\omega| \Delta t\right)\right]^{1/3} \end{equation} \noindent All singular evolutions have particle horizons because $a\sim \Delta t^{1/3}$ as $\Delta t\to 0$. For other values of $\gamma$ and $k$, there is also a critical value of the cosmological constant $\Lambda_0=-3\zeta^2/(4\gamma^2)$, so that for $\Lambda>\Lambda_0$ an asymptotically de Sitter evolution occurs for two values of $H$ \begin{equation} \label{18} H_\pm=\frac{1}{2\gamma}\left[\zeta\pm\left(\zeta^2+\frac{4}{3}\gamma^2\Lambda \right)^{1/2}\right] \end{equation} \noindent We verify that the behavior $a\sim\exp(H_+t)$ is asymptotically stable by means of the Liapunov function $L_+=(H-H_+)^2$, which satisfies $L_+'<0$ for large times when $H>0$. Nonlinear viscous effects has been shown to arise as a phenomenological description of the effect of particle creation in the early universe \cite{Ver}, and cosmological models with this kind of fluids has been considered in \cite{Nov} and \cite{Rom}. Here, we assume that $\zeta=\alpha H$, with a constant $\alpha$ such that $\gamma>\alpha>0$, and we consider only expanding evolutions. Following the same procedure as before, we arrive at the equation \begin{equation} \label{20} \ddot v+M v^m+N v^n-\frac{\gamma\Lambda}{2\nu}v=0 \end{equation} \noindent where now $\nu=2/[3(\gamma-\alpha)]$, $M=(3/8)(\gamma-\alpha)(2-\gamma)$, $m=1-4/(\gamma-\alpha)$, $N=(3/4)(\gamma-\alpha)(3\gamma-2)k$ and $n=1-4/[3(\gamma-\alpha)]$. As this case behaves as a conservative mechanical system, we may obtain (at least formally) its general solution: \begin{equation} \label{21} \Delta t=\int\frac{dv}{\sqrt{2\left[E-V(v)\right]}} \end{equation} \noindent where $E$ is an integration constant and \begin{equation} \label{22} V(v)=\frac{M}{m+1}v^{m+1}+\frac{N}{n+1}v^{n+1}-\frac{\gamma\Lambda}{4\nu}v^2 \qquad \gamma-\alpha\neq\frac{2}{3} \end{equation} \begin{equation} \label{23} V(v)=\frac{M}{m+1}v^{m+1}+N\ln v-\frac{\gamma\Lambda}{4\nu}v^2 \qquad \gamma-\alpha=\frac{2}{3} \end{equation} A qualitative analysis of (\ref{21}) shows that expanding evolutions may begin either at a singularity or a bounce. Both bounded and unbounded singular solutions occur, but only for $\Lambda<0$ are there bouncing solutions that reach a maximum. We quote two cases for which the equation (\ref{20}) becomes linear, so that we may obtain its general solution in closed form (for simplicity we take $\Lambda=0$ ). The first one is $\gamma=2$ and $k=0$ \begin{equation} \label{24} a(t)=\left(A\Delta t\right)^\nu \end{equation} \begin{equation} \label{25} \Delta\phi(t)=\frac{C\Delta t^{1-3\nu}}{A^{3\nu}(1-3\nu)} \end{equation} \begin{equation} \label{26} \rho(t)=\frac{3\nu^2}{\Delta t^2}-\frac{C^2}{2(A\Delta t)^{6\nu}} \end{equation} \noindent The other solution is for $\gamma=2$, $\alpha=2/3$ and $k\neq0$. We find \begin{equation} \label{27} a(t)=\left|A\Delta t-2k\Delta t^2\right|^{1/2} \end{equation} \begin{equation} \label{28} \Delta\phi(t)=-\frac{2C}{A^2}\frac{A-4k\Delta t} {\left|A\Delta t-2k\Delta t^2\right|^{1/2}} \end{equation} \begin{equation} \label{29} \rho(t)=\frac{3}{4}\frac{(A-4k\Delta t)^2} {\left|A\Delta t-2k\Delta t^2\right|^2}-\frac{1}{2}\frac{C^2} {\left|A\Delta t-2k\Delta t^2\right|^3}+\frac{3k} {\left|A\Delta t-2k\Delta t^2\right|} \end{equation} We find also solutions of the form $a=a_0 \Delta t^\sigma$ for special values of $\alpha$: a. $\gamma=2$, $\sigma=1$, $\alpha=(4/3)(1+k/a_0^2)$; b. $k=0$, $\sigma=1/3$, $\alpha=\gamma+3(2-\alpha)C^2/(2a_0^2)-2$. \section{Conclusions} We have found exact solutions of the Einstein equations with a free scalar field and a viscous fluid source, in a homogeneous isotropic metric. We have considered cosmological models with two forms of the bulk viscosity pressure: linear, with a constant coefficient, and cuadratic in the Hubble variable. For the first case the picture is similar to the case without scalar field and we show that the de Sitter evolution is asymptotically stable. Also, we find that the physical requirement of positivity of the energy density of the fluid makes some of these solutions geodesically incomplete. For the case of a nonlinear viscous pressure we reduce the equations of this model to that of a conservative mechanical system. Thus, we are able to give its general solution in implicit form, and besides we show several cases for which an explicit solution arise. Both, singular and bouncing solutions arise, and we find cases in which the scale factor reach a maximum or grow without bound. In the latter case, when $\Lambda=0$, the scale factor has a power law behavior for large times. Dissipative effects like particle production may have been important in the early universe, for instance in the reheating period at the end of the inflationary era. It is generally assumed that this stage of accelerated evolution was driven by a self-interacting scalar field. Thus, we consider of interest to investigate further the models of this paper, taking into account also an interaction potential. \newpage
\subsection*{To appear in Physical Review B.} \subsection*{cond-mat/9508140} \hspace*{0.5cm} \section{ Introduction } In our recent paper \cite{anyon} we have pointed out the possibility that vortices in superconducting films might be anyons. There is a classic paper by Haldane and Wu \cite{hawu}, which demonstrates that vortices in superfluid helium layers are not anyons because there is no well defined dilute limit. We have reanalysed this question from the point of view of the phenomenological time-dependent Ginzburg-Landau models. It was shown that there are situations when one can go around the argument in \cite{hawu}. A classic example are vortices in the Ginzburg-Landau models of the fractional quantum Hall effect \cite{wl,glhe}. Chern-Simons interaction makes them well localised objects and the dilute limit can be defined. The main effect of the CS gauge field is to remove the divergent gradient energies present in the global model. These divegencies are also removed in the Ginzburg-Landau model for superconductors. We have considered a gauged nonlinear Schrodinger equation as a minimal version of time-dependent Ginzburg-Landau model. To model the structure of the vortex core we assumed it was filled with normal fluid so as to make the whole structure locally charge-neutral. The strenght of the statistical interaction was proportional to the net deficit of the superfluid replaced inside the core by the normal phase. It is argued \cite{baps} that the time evolution of the condensate is much faster then that of the normal fluid. Thus in the static case the core is filled with the normal fluid but when the vortex moves fast enough the normal fluid can not adjust itself and does not follow the vortex motion. It is a crude approximation \cite{stoof} and it does not invalidate our results. The statistical interaction shows up in the adiabatic approximation which is quite an oposite limit. In this limit we can assume the nonuniformity of the normal fluid does follow the vortex motion. The effective Lagrangian for vortex motion can be arranged term by term according to the powers of vortex velocity. The statistical interaction together with the term responsible for the Magnus force \cite{at} are the first terms in this effective Lagrangian. The distortion of the normal fluid distribution from the static one, which is at least of the first order in velocities, can contribute but only to the term quadratic in velocities which is the next to leading term. Thus at least for slow motion as compared to characteristic velocities, when expansion in powers of velocities is justified, the Magnus force and statistical interaction are not altered. One can rewrite the local BCS model in terms of the gap-function field \cite{abts,stoof,aatz} but the effective theory appears to be nonlocal - it contains derivatives of arbitrary order. In this way one goes from description in terms of electronic degrees of freedom to description in terms of Cooper pairs. The latter is well suited for the bulk of the superconductor. However in the core of the vortex we can expect some decoherence, which phenomenologically might be decribed by normal fluid. Whenever degrees of freedom other than those of Cooper pairs come into play the description in terms of Cooper pairs may happen to be irrelevant. One way of dealing with the problem is to truncate the effective theory on the lowest order in derivatives and introduce more or less explicitly something like a normal component. This approach is limited by the poor knowledge about the nature of the normal fluid. Another approach is to take as many orders in derivatives in the effective theory as possible to describe also the normal fluid in terms of the gap function. The problem is that in practice one would have to cut the expansion at certain order and there is no warranty that such a cut theory would be self-consistent. To go around these problems we will derive the statistical interaction in the framework of the Bogolubov-de Gennes formalism for pure samples at zero temperature. We will work in the quasi two-dimensional regime of long parallel vortices. In the case of superconducting layers thicker than $100\AA$ the penetration lenght $\Lambda$ is still very close to the penetration lenght $\lambda$ in the bulk superconductor. In this regime we expect modifications to be rather quantitative in nature then qualitative. An important first step was done by Gaitan \cite{gaitan} in his derivation of the Berry phase responsible for the Magnus force. In this paper we are going to reanalize his derivation and then to extend the method to the case of two well separated vortices. In the microscopic theory the language of the normal and superfluid is not very fruitful. The distinction of the states below the Fermi surface into localised bound states and scattering states appears to be more natural. The distinction does not influence the value of the Magnus force but it is crucial for the statistical interaction. The scattering states are common to all vortices while bound states can be identified with particular ones. \section{ Preliminaries on vortex solution in BCS theory } The problem of the vortex solution in the BCS theory can be conveniently posed within the Bogolubov-de Gennes formalism \cite{dg}. It has not been completely solved although some qualitative fictures of the solution are known \cite{bardeen}. We will restrict here to listing the basic ingredients of the formalism. The Bogolubov equation, defined in the Nambu spinor space \cite{nambu}, is \begin{equation}\label{2.10} (E_{n}-\hat{H}_{BOG}) \left(\begin{array}{c} u_{n} \\ v_{n} \end{array}\right)=0 \;\;, \end{equation} where $\hat{H}_{BOG}$ is the Bogolubov hamiltonian \begin{equation}\label{2.20} \hat{H}_{BOG}= \left[\begin{array}{cc} -\frac{1}{2}(\mbox{\boldmath $\nabla$}-ie\mbox{\boldmath $A$})^{2}-E_{F} & \Delta(\mbox{\boldmath $x$}) \\ \Delta^{\star}(\mbox{\boldmath $x$}) & \frac{1}{2}(\mbox{\boldmath $\nabla$}+ie\mbox{\boldmath $A$})^{2}+E_{F} \end{array} \right] \;\;. \end{equation} $E_{F}$ is the Fermi energy. $\Delta(\mbox{\boldmath $x$})=\Delta_{0}(r)\exp(-i\theta)$ denotes the gap function of vortex solution. $\Delta_{0}(r)$ interpolates between $0$ at the origin and a constant, which we will call $\sqrt{\rho_{0}}$, at infinity. There are both positive and negative energy solutions. If $(u_{n},v_{n})$ is an eigenstate with energy $E_{n}>0$, then $(-v^{\star}_{n},u^{\star}_{n})$ is a solution with energy $-E_{n}<0$. The equations (\ref{2.10}) have to be supplemented by a self-consistency condition \begin{equation}\label{2.25} \Delta(\mbox{\boldmath $x$})= g\sum_{n}u_{n}(\mbox{\boldmath $x$})v_{n}^{\star}(\mbox{\boldmath $x$}) \;\;, \end{equation} where $g$ is the BCS coupling constant, together with Maxwell equations determining the vector potential. The field operator for Nambu quasiparticles can be expanded in terms of the solutions of Eq.(\ref{2.10}). \begin{equation}\label{2.30} \Psi( \mbox{\boldmath $x$} )=\left(\begin{array}{c} \psi_{\uparrow}(\mbox{\boldmath $x$}) \\ \psi_{\downarrow}^{\dagger}(\mbox{\boldmath $x$}) \end{array}\right)= \sum_{n} \left[\begin{array}{c} \gamma_{n\uparrow} \left(\begin{array}{c} u_{n}(\mbox{\boldmath $x$}) \\ v_{n}(\mbox{\boldmath $x$}) \end{array}\right) + \gamma_{n\downarrow}^{\dagger} \left(\begin{array}{c} -v^{\star}_{n}(\mbox{\boldmath $x$}) \\ u^{\star}_{n}(\mbox{\boldmath $x$}) \end{array}\right) \end{array}\right] \;\;. \end{equation} The negative energy states are occupied in the BCS ground state \begin{equation}\label{2.40} \mid BCS > = \prod_{n} \gamma_{n\downarrow} \mid 0 > \;\;. \end{equation} The eigenstates satisfy the following orthogonality \begin{eqnarray}\label{2.50} &&\int d^{3}x\; [u_{n}(\mbox{\boldmath $x$})u^{\star}_{m}(\mbox{\boldmath $x$}) +v_{n}(\mbox{\boldmath $x$})v^{\star}_{m}(\mbox{\boldmath $x$})]= \delta_{nm} \;\;,\nonumber\\ &&\int d^{3}x\; [u_{n}(\mbox{\boldmath $x$})v_{m}(\mbox{\boldmath $x$}) -v_{n}(\mbox{\boldmath $x$})u_{m}(\mbox{\boldmath $x$})]=0 \end{eqnarray} and completeness relations \begin{eqnarray}\label{2.60} &&\sum_{n}[u_{n}(\mbox{\boldmath $x$}) u_{n}^{\star}(\mbox{\boldmath $x$}^{\prime}) +v_{n}^{\star}(\mbox{\boldmath $x$}) v_{n}(\mbox{\boldmath $x$}^{\prime})] =\delta(\mbox{\boldmath $x$}-\mbox{\boldmath $x$}^{\prime}) \;\;,\nonumber\\ &&\sum_{n}[u_{n}(\mbox{\boldmath $x$}) v_{n}^{\star}(\mbox{\boldmath $x$}^{\prime}) -v_{n}^{\star}(\mbox{\boldmath $x$}) v_{n}(\mbox{\boldmath $x$}^{\prime})]=0 \;\;. \end{eqnarray} With the help of these relations the creation and annihilation operators can be expressed as \begin{eqnarray}\label{2.70} &&\gamma_{n\downarrow}=\int d^{3}x\; [-\psi^{\dagger}_{\uparrow}(\mbox{\boldmath $x$}) v_{n}^{\star}(\mbox{\boldmath $x$}) +\psi_{\downarrow}(\mbox{\boldmath $x$}) u^{\star}_{n}(\mbox{\boldmath $x$})] \;\;, \nonumber\\ &&\gamma_{n\uparrow}^{\dagger}=\int d^{3}x\; [\psi^{\dagger}_{\uparrow}(\mbox{\boldmath $x$}) u_{n}^{\star}(\mbox{\boldmath $x$}) +\psi_{\downarrow}(\mbox{\boldmath $x$}) v_{n}(\mbox{\boldmath $x$})] \;\;. \end{eqnarray} Adiabatic vortex motion \cite{gaitan} gives rise to a Berry phase in the solutions of Eq.(\ref{2.10}), $(u_{n},v_{n})\rightarrow \exp[i\phi_{n}](u_{n},v_{n})$. These Berry phases sum up to the total Berry phase picked up by the ground state $\mid BCS >\rightarrow \exp[i\Gamma] \mid BCS >$ which with the help of Eqs.(\ref{2.70},\ref{2.40}) can be established to be \begin{equation}\label{2.80} \Gamma=-\sum_{n}\phi_{n} \;\;. \end{equation} To persue some of the questions we need more detailed knowledge about the eigenstates. The axially symmetric ansatz takes the form \begin{equation}\label{2.90} \chi_{n}(\mbox{\boldmath $x$})=\left ( \begin{array}{c} u_{n}(\mbox{\boldmath $x$}) \\ v_{n}(\mbox{\boldmath $x$}) \end{array} \right )= e^{ik_{z}z} e^{i(\mu-\frac{1}{2}\sigma_{z})\theta} f_{n}(r) \;\;, \end{equation} where $\hbar k_{z}$ is the $z$-component of the momentum. $\mu$ must be a half-integer for the expression to be single-valued. The functions $f_{n}(r)$ have been investigated in \cite{bardeen} with the help of WKB approximation. We will quote some more detailed results in the following. \section{ Origin of the Magnus force } Now we are going to rederive the Berry phase responsible for the Magnus force following the argument of Gaitan \cite{gaitan}. In comparison with \cite{gaitan} we clarify some points and remove some unnecessary assumptions. The general form of the Berry phase in two dimensions is \begin{equation}\label{3.10} \phi_{n}=i\int dt\; \int d^{2}x\; \chi_{n}^{\dagger} (\frac{d}{dt}+i\frac{e}{\hbar}A_{0}) \chi_{n} \;\;. \end{equation} The time derivative is understood as a total derivative with respect to slow degrees of freedom. For an adiabatic motion of a single vortex the derivative has to be replaced by $\dot{\mbox{\boldmath $r$}}_{0}\mbox{\boldmath $\nabla$} _{\mbox{\boldmath $r_{0}$}}$, where $\mbox{\boldmath $ r_{0} $}$ is a position of the vortex singularity. The scalar potential vanishes for the vortex solution so we will skip the second term in what follows. With the axially symmetric ansatz (\ref{2.90}) the phase becomes \begin{equation}\label{3.20} \phi_{n}=\int dt\;\int d^{2}x\; \dot{ \mbox{\boldmath $r$} }_{0} [ (\chi^{\dagger}_{n}(-\mu+\frac{1}{2}\sigma_{z})\chi_{n}) \mbox{\boldmath $ \nabla $}_{\mbox{\boldmath $r_{0}$}} \theta+ f^{\dagger}_{n}(r) \mbox{\boldmath $ \nabla $}_{\mbox{\boldmath $r_{0}$}} f_{n}(r)] \end{equation} The second term vanishes by symmetry arguments. The contribution of the first term to the total Berry phase is \begin{equation}\label{3.30} \Gamma=-\sum_{n}\phi_{n}=-\int dt\;\int d^{2}x\; ( \dot{ \mbox{\boldmath $r$} }_{0} \mbox{\boldmath $ \nabla $} _{ \mbox{\boldmath $ r_{0} $} }\theta) S(r) \end{equation} where \begin{equation}\label{3.40} S(r)=\sum_{n}[\mid u_{n}(r)\mid^{2}(-\mu+\frac{1}{2})+ \mid v_{n}(r)\mid^{2}(-\mu-\frac{1}{2})] \;\;. \end{equation} $\hbar S$ is minus the z-component of the canonical angular momentum density. It is not a gauge-invariant integral of motion. $S$ is the expectation value density of the operator $-i\hbar\frac{\partial}{\partial\theta}$ instead of the gauge-invariant $-i\hbar\frac{\partial}{\partial\theta}+\sigma_{z}A_{\theta}$. $S(0)=0$ because either $u_{n}$ ($v_{n}$) or the factor $(-\mu\stackrel{+}{-}\frac{1}{2})$ vanishes at the origin. To find out its asymptotic behavior at infinity we would need a much more detailed knowledge about the solutions. We go around this problem by resorting to the effective theory which is equivalent to the microscopic formalism. The general term linear in the covariant time derivative reads \begin{equation}\label{3.50} \int dt\;\int d^{2}x\; i[\Delta^{\star}(\hbar\partial_{t}+2ieA_{0})\Delta-c.c.] [G(\Delta^{\star}\Delta)+spatial\;derivative\;terms]\;\;. \end{equation} By "$spatial\;derivative\;terms$" we mean terms which are of at least first order in the covariant spatial derivatives. $G$ is a function of $\Delta^{\star}\Delta$ only which tends to $\rho_{s}/\rho_{0}$ as the gap function $\Delta$ approaches its asymptotic equilibrium value. $\rho_{s}$ is the equilibrium Cooper pairs' density. Let us consider the adiabatic rotation of the vortex solution $\Delta=\Delta_{0}(r)e^{-i\theta}$ around its axis, $\theta\rightarrow\theta-\omega t$. The action picks up a term (to lowest order in $\omega$) \begin{equation}\label{3.60} \omega \int dt\;\int d^{2}x\; \{-\hbar\Delta^{\star}\Delta [G(\Delta^{\star}\Delta)+spatial\;derivative\;terms]\} \;\;. \end{equation} The spatial integral is just the total angular momentum. For large $r$, where $A_{0}$ tends to zero, the density of this angular momentum is, by gauge invariance (\ref{3.50}), equal to $\hbar$ times minus the bulk Cooper pairs' density $\hbar S\approx-\hbar\rho_{s}=\hbar\lim_{r\rightarrow\infty} \frac{\delta W}{\delta (2eA_{0})}$, where $W$ is the effective action and $\rho_{s}>0$. We have to stress that we make use of the effective theory only very far from the vortex core where it should be equivalent to the microscopic treatement. In particular in the distant asymptotic region there is no contribution from unpaired bound states which can not be described in terms of Cooper pairs. We have all we need to calculate the Berry phase. Let us expand the integrand in Eq.(\ref{3.30}) around the vortex position $\mbox{\boldmath $r_{0}$}=(X,Y)$ close to the origin, $(X,Y)=(0,0)$, \begin{eqnarray}\label{3.70} &&S=S(r)-S^{\prime}(r)[X\cos\theta+Y\sin\theta] +O(r_{0}^{2}) \;\;,\nonumber\\ &&\dot{ \mbox{\boldmath $r$} }_{0}\mbox{\boldmath $\nabla$} _{\mbox{\boldmath $r_{0}$}}\theta= \dot{X}(\frac{\sin\theta}{r}+\frac{X\sin 2\theta-Y\cos 2\theta}{r^{2}})+ \dot{Y}(-\frac{\cos\theta}{r}-\frac{X\cos 2\theta+Y\sin 2\theta}{r^{2}})+ O(r_{0}^{2}) \;\;. \end{eqnarray} A straightforward integration yields \begin{equation}\label{3.80} \Gamma=-\pi [S(\infty)-S(0)] \int dt\;\varepsilon_{kl}X^{k}\dot{X}^{l} +O(r_{0}^{2}) \;\;. \end{equation} The expression $O(r_{0}^{2})$ does vanish. We are considering single vortex in absence of any driven current. Such a system is translationally invariant and isotropic. The first term on the R.H.S. of Eq.(\ref{3.80}) is already the most general term linear in velocity which is, up to a total time derivative, translationally invariant and isotropic. Thus we do not need to consider finite $\mbox{\boldmath $r_{0}$}$ to obtain a generally valid expression. The phase (\ref{3.80}) is remarkably simple to evaluate. For a vortex with winding number $-1$ it reads \begin{equation}\label{3.90} \Gamma=\pi\rho_{s}\int dt\; \varepsilon_{kl}X^{k}\dot{X}^{l} \;\;. \end{equation} From our derivation of the Magnus force it is clear that the Wess-Zumino term (in the gauge $A_{0}=0$) \begin{equation}\label{3.100} \int dt\;\int d^{2}x [\rho_{s}\partial_{t}\theta] \;\;, \end{equation} with $\rho_{s}=const$, does not make much sense as it stands. $\rho_{s}$ is the same at the origin as at infinity so the Magnus force vanishes. The formula (\ref{3.100}) is to be understood with an implicit assumption that a small area around the phase singularity is excluded from the spatial integration. In other words $\rho_{s}$ must be put equal to $0$ in this area. It is not difficult to realise, by performing radial integration first and then integration over the angle around the singularity, that the way of regularisation does not matter. In particular it does not need to be rotationally symmetric. The only factors that determine the Magnus force are the two limit values of $\rho_{s}$. Thus vortices in a condensate will always feel the Magnus force. It is not the case for say Jackiw-Pi solitons \cite{jp}, where $\rho_{s}$ is zero both at the origin and at infinity. \section{ Mutual statistical interaction of vortices } Let us consider two vortices: "1" at the origin and "2" very far apart at $\mbox{\boldmath $R$}(t)$. It is important to realise that the eigenstates of the Bogolubov hamiltonian can be divided into common scattering states, which we will still denote by just $u_{n},v_{n}$, and bound states which can be identified with a given vortex $u^{(1,2)}_{n},v^{(1,2)}_{n}$. Vortices are very distant so there is no overlap between their localised bound states. The bound states of the stationary vortex "1" feel what is going on around them through the pair potential $\Delta(t,\mbox{\boldmath $x$})$ inside and around the core. Vortices are well localised so a fairly good approximation to a two-vortex gap function is the product ansatz \begin{equation}\label{4.10} \sqrt{\rho_{0}}\Delta(t,\mbox{\boldmath $x$}) =\Delta_{v}( \mbox{\boldmath $x$} ) \Delta_{v}[ \mbox{\boldmath $x-R$}(t) ] \;\;, \end{equation} where $\Delta_{v}(\mbox{\boldmath $x$})=\Delta_{0}(r)\exp(-i\theta)$ denotes the gap function of a single vortex centered at the origin. Close to $\mbox{\boldmath $x$}=0$ this expression can be further simplified \begin{equation}\label{4.20} \Delta(t,\mbox{\boldmath $x$})=\Delta_{v}(\mbox{\boldmath $x$}) e^{ -i\theta[\mbox{\boldmath $x-R$}(t)] } \;\;, \end{equation} Thus the bound states of the static vortex have to be modified as \begin{equation}\label{4.30} \chi_{n}^{(1)}[\mbox{\boldmath $x,R$}(t)] = e^{-i\frac{\sigma_{z}}{2}\theta[\mbox{\boldmath $x-R$}(t)]} \chi_{n}(\mbox{\boldmath $x$}) \;\;. \end{equation} Their contribution to the Berry phase is \begin{eqnarray}\label{4.40} -\int dt\;\int d^{2}x\; \{ \dot{ \mbox{\boldmath $R$} } \mbox{\boldmath $\nabla$} _{\mbox{\boldmath $R$}}\theta[\mbox{\boldmath $x-R(t)$}] \} \sum_{bound\;st.}(\frac{1}{2}\mid u_{n}\mid^{2}- \frac{1}{2}\mid v_{n}\mid^{2})\approx \nonumber\\ \{\frac{1}{2}\int d^{2}x\; \sum_{bound\;st.}(\mid u_{n}\mid^{2}-\mid v_{n}\mid^{2})\} \int dt\; \dot{ \mbox{\boldmath $R$} }\mbox{\boldmath $\nabla$} _{\mbox{\boldmath $R$}}\theta[\mbox{\boldmath $R(t)$}]\;\;, \end{eqnarray} where the approximate equality is valid for small $\frac{r_{c}}{R}$, where $r_{c}$ is a radius of the core. The equality $\mid u_{n}\mid^{2}=\mid v_{n}\mid^{2}$ holds for the bound states, at least up to the WKB approximation \cite{bardeen}, so their contribution to the Berry phase vanishes. Now as the vortex "2" moves its bound states follow its trajectory $\chi^{(2)}[\mbox{\boldmath $r-R$}(t)]$, similarly as in the single vortex case considered in the previous section. In addition, as an effect due to the vortex "1" (\ref{4.10}), their components perform the relative phase rotation \begin{equation}\label{4.50} \chi_{n}^{(2)}[\mbox{\boldmath $x,R$}(t)] \approx e^{-i\frac{\sigma_{z}}{2}\theta[\mbox{\boldmath $R$}(t)]} \chi_{n}[\mbox{\boldmath $x-R(t)$}] \;\;. \end{equation} The contribution from this relative phase rotation is once again zero. The bound states contribute but only to the Magnus term (\ref{3.90}) just as in the single vortex case. The bound states do not give rise to any new effects so let us consider scattering states common to both vortices. In the vortex core region the asymptotes of the scattering states must be close to those of the scattering states for a single vortex but the phase has to be replaced by the asymptote of the phase in the product ansatz (\ref{4.20}). Close to the origin \begin{equation}\label{4.60} \chi_{n}[\mbox{\boldmath $x,R$}(t)]\approx e^{ik_{z}z} e^{i(\mu-\frac{\sigma_{z}}{2}) \{\theta(\mbox{\boldmath $x$})+\theta[\mbox{\boldmath $x-R$}(t)]\}} f_{n}(r) \;\;. \end{equation} The contribution from around the stationary vortex is \begin{equation}\label{4.70} -\int dt\;\int d^{2}x\; \dot{\mbox{\boldmath $R$}}\mbox{\boldmath $\nabla$}_{\mbox{\boldmath $R$}} \theta[\mbox{\boldmath $x-R$}(t)] \bar{S}(r) \;\;, \end{equation} where $\bar{S}$ is a part of the canonical angular momentum due to the scattering states \begin{equation}\label{4.80} \bar{S}(r)=\sum_{scatt.\;st.}[\mid u_{n}(r)\mid^{2}(-\mu+\frac{1}{2})+ \mid v_{n}(r)\mid^{2}(-\mu-\frac{1}{2})] \;\;. \end{equation} Far from the core $\bar{S}\approx S\approx -\rho_{s}$. If $\bar{S}$ were equal to $-\rho_{s}$ also in the core, the contribution to the Berry phase from (\ref{4.70}) would be just the same as to the Magnus term. Thus we are interested only in the effects due to deviations of $\bar{S}$ from its asymptotic value $-\rho_{s}$. Inside the core $\mid u_{n}\mid^{2}$ and $\mid v_{n}\mid^{2}$ are changed by a factor which is $>1$ for the states with $\mu$ negative and $<1$ for $\mu$ positive \cite{bardeen}. The net deviation $\delta\bar{S}(r)=\bar{S}(r)+\rho_{s}$ is positive. At the very origin $\delta\bar{S}(0)=\rho_{s}$. The total change in the Berry phase is twice that in (\ref{4.70}), as there are two vortices, and amounts to \begin{equation}\label{4.90} \delta\Gamma=[2\int d^{2}x\;\delta\bar{S}(r)] \int dt\; \dot{\mbox{\boldmath $R$}}\mbox{\boldmath $\nabla$} _{\mbox{\boldmath $R$}}\theta(\mbox{\boldmath $R$})\;\;. \end{equation} Thus the total Berry phase for a dilute vortex system is \begin{equation}\label{4.100} \Gamma=\int dt\; [ -\pi\rho_{s}\sum_{p} n_{p}\varepsilon^{kl}X^{k}_{(p)}\dot{X}^{l}_{(p)}+ \alpha\sum_{p<q}n_{p}n_{q}\frac{d}{dt}\Theta_{(p,q)}] \;\;, \end{equation} where the indices $p,q$ run over vortices, $n$'s are their winding numbers, $\Theta_{(p,q)}$ is the angle between the $p$-th and $q$-th vortex and the numerical factor $\alpha$ can be read from Eq.(\ref{4.90}) \begin{equation}\label{4.110} \alpha=2 \int d^{2}x\; \delta\bar{S}(r) \;>\; 0 \;\;. \end{equation} $\alpha$ is roughly the number of electrons inside the core and as such it can range from $\sim 1$ for high $T_{c}$ superconductors to $\sim 10^{5}$ for some conventional type $II$ superconductors. \section{ Vortex statistics within variational wave-function approach } Once we have derived statistical interaction in the microscopic setting it may be worthwhile to reanalise some earlier approaches to similar problems. In the paper by Ao and Thouless \cite{at} the Magnus force was derived with the help of the variational many-electron vortex wave-function \begin{equation}\label{at.10} \psi_{v}[z]=\exp[\frac{i}{2}\sum_{k}\theta(z_{k}-z_{0})]\psi_{0}[z] \;\;, \end{equation} where $z_{0}$ is a complex vortex position, $z_{k}$'s are positions of electrons and $\psi_{0}[z]$ is an antisymmetric variational function. The phase factors in the wave-function are determined by the demand of correct electronic quantumstatistics and by topological properties. There are variational profile functions in $\psi_{0}$ which can not be established without dynamical considerations. One can consider an adiabatic vortex motion along some trajectory and calculate the Berry phase picked up by the wave-function. This Berry phase coinsides with Eq.(\ref{3.90}). For a closed path the Berry phase is proportional to the number of electrons enclosed by the trajectory. This setting is convenient to analise what is the dependence of the Magnus force on impurities \cite{at}. An impurity can be viewed as an attractive potential which traps some of electrons in localised bound states. The trapped electrons disappear from the ansatz (\ref{at.10}). The Berry phase is still proportional to the area enclosed by the trajectory but this time the area should not be multiplied by the total density of electrons but rather by the total density minus the density of electrons trapped by impurities. Impurities lower the value of the Magnus force. Now let us consider the effect of an exchange of two vortices. More precisely, let us fix the position of one vortex and consider another distant vortex moving around it. One could argue there is no special effect because the net charge of any vortex must be zero. Provided the trajectory is large enough, there is no change in the number of enclosed electrons due to the enclosed vortex. The last sentence is certainly true but the example with impurities tought us that it is not the total number of electrons that really matters but rather the number of electrons in the coherent state described by the wave-function (\ref{at.10}). We know from the discussion in the previous sections that inside vortex core the scattering or continuum states are replaced by bound states. Thus vortex can be viewed as a kind of impurity, which traps some of electrons into localised bound states with energies within the energy gap band. The localised electrons are removed from the wave-function (\ref{at.10}). There is an additional Berry phase proportional to the number of vortices enclosed by the trajectory. Each enclosed vortex contributes a term proportional to the number of electrons trapped inside its core. We can consider a path for a chosen vortex in a more or less uniform distribution of vortices. If we neglect possible intervortex correlation effects, the background vortices could be regarded as uniform distribution of impurities lowering the density of electrons in the coherent state (\ref{at.10}). In this mean-field approximation the Magnus force acting on a choosen vortex is lowered by the presence of another vortices. This approximation is nothing else but the delocalisation procedure so often applied to anyonic systems. The Magnus force can be interpreted as Lorenz force due to interaction of effectively charged vortices with some uniform effective magnetic field. The statistical interaction can be seen as Aharonov-Bohm effect due to the fluxes attached to vortices. In the mean-field approximation the fluxes, which are opposite to the external flux, are delocalised and they lower the net uniform flux. In the same way the real impurities can be interpreted as localised fluxes, opposite to the external field, randomly distributed over the plane. If the M-F approximation appears to work for real impurities, it will also work for vortices. \section{ Hall angle and vortex density } The fact that the value of the Magnus force can be lowered with increasing density of vortices can, in principle, be observable in Hall experiments \cite{hl}. The vortex density should increase and the M-F Magnus force should decrease with increasing real magnetic field. This should manifest itself in the changes of the measured Hall angle. Vortex equation of motion takes the form \cite{ah} \begin{equation}\label{at.20} m_{eff}\ddot{\mbox{\boldmath $r$}}= \frac{\rho_{s}hd}{2} ( \dot{\mbox{\boldmath $r$}} - \mbox{\boldmath $v_{s}$} )\times \hat{\mbox{\boldmath $z$}} -\eta d \dot{\mbox{\boldmath $r$}} + \mbox{\boldmath $F_{pin}$} + \mbox{\boldmath $f$} \;\;, \end{equation} where $m_{eff}$ is a small effective vortex mass, $\eta$ is a vortex viscosity, $\mbox{\boldmath $F_{pin}$}$ is a pinning force, $\mbox{\boldmath $f$}$ is a fluctuating force, $d$ is a sample thickness and $\mbox{\boldmath $v_{s}$}$ is a driven uniform superfluid velocity. When we neglect pinning and average over fluctuations the stationary state motion will be determined by the equation \begin{equation}\label{at.30} ( \dot{\mbox{\boldmath $r$}} - \mbox{\boldmath $v_{s}$} )\times \hat{\mbox{\boldmath $z$}} =\tan(\theta_{H})\dot{\mbox{\boldmath $r$}} \;\;, \end{equation} where $\tan(\theta_{H})=\frac{2\eta}{\rho_{s}h}$. The solution is \begin{equation}\label{at.40} \dot{\mbox{\boldmath $r$}}= \frac{ \mbox{\boldmath $v_{s}$} + (\hat{\mbox{\boldmath $z$}}\times\mbox{\boldmath $v_{s}$}) \tan(\theta_{H}) }{1+\tan^{2}(\theta_{H}) } \;\;. \end{equation} $\theta_{H}$ is the angle between the superfluid velocity $\mbox{\boldmath $v_{s}$}$ and the stationary vortex velocity. The angle is the larger the weaker is the Magnus force. If the effective Magnus force is lowered with increased vortex density the angle should also grow with external magnetic field, which drives the rise in vortex density. The changes of the Magnus force due to changes in vortex density should be the more rapid the larger is the number of electrons trapped in the vortex core. For this reason we would recommend experiments on mildly type $II$ conventional superconductors with large vortex cores (large correlation length $\xi$). Rather strong viscosity should be prefered for the angle to be more sensitive to the strengh of the Magnus force. The sample should be pure of pinning centers to avoid obscure pinning effects. \section{ Vortices' fractional quantum Hall effect } To summarise our knowledge about the dynamics of planar vortices, let us write down an effective Lagrangian for diluted vortices with topological charge $-1$ \begin{equation}\label{5.10} L_{eff}=\sum_{p} [\frac{1}{2}m_{eff}\dot{\mbox{\boldmath $X$}}_{p} \dot{\mbox{\boldmath $X$}}_{p} +\hbar\pi\rho_{s} \mbox{\boldmath $X_{p}$} \times \dot{ \mbox{\boldmath $X$} }_{p} ] +\sum_{p<q}[\hbar\alpha\frac{d}{dt}\Theta_{(p,q)} -V_{eff}(\mid\mbox{\boldmath $X_{p}-X_{q}$}\mid)] \;\;. \end{equation} The indices $p,q$ run over vortices. $m_{eff}$ is the effective vortex mass. It is usually estimated to be around $10^{8}m_{e}/m=10^{-2}m_{e}/\AA$. The second term in Eq.(\ref{5.10}) decribes interaction of vortices with effective magnetic field. We stress that this field has nothing to do with the real magnetic field $B_{ext}$, which in this case is just a device to drive the changes of vortex density. If we assumed the density of electrons to be $\sim 10^{30} m^{-3}=1\AA^{-3}$ the effective magnetic field defined by $\frac{eB_{eff}}{2}=\pi\hbar\rho_{s}$ would turn out to be $\sim 10^{6} T/\AA$. When compared with the effective vortex mass per $1\AA$ the magnetic field turns out to be incredibly strong. Its effect on a vortex should be the same as that of the $10^{8}T$ magnetic field on an electron. Vortices can be expected to be confined to the lowest Landau level (LLL). Now let us consider a single vortex at $z=X_{1}+iX_{2}$. What is the magnetic lenght $l$ which determines the size of the LLL wavepacket \begin{equation}\label{5.20} \psi_{0}(z,\bar{z})= \frac{1}{\sqrt{2\pi l^{2}}}e^{-\frac{z\bar{z}}{4l^{2}} }\;\;? \end{equation} The magnetic lenght is determined by the strength of the exactly known Magnus force $l^{2}=\frac{1}{2\pi\rho_{s}}$. The area over which the center of such a vortex fluctuates can be estimated to be $\frac{2}{\rho_{s}}$, which is of the same order as the area per one electron. This effect is significant for extremely type $II$ superconductors where the core is very thin. Now we are prepared to address the question of the fractional Hall effect. Vortices are anyons with a statistical parameter $\alpha$. A nontrivial statistics would be sufficient to prevent them from overlapping if the third term in (\ref{5.10}) were not regularised at short distances and replaced by mutual charge-flux interaction \cite{kld}. Fortunately we also have an effective short range mutual repulsion $V_{eff}$ which at low temperatures may be sufficient to keep vortices at a distance. On the other hand the potential is very weak as compared to the Landau energy so the mixing with higher LL's should be in any case negligible. Following Laughlin \cite{laughlin}, the trial wave-function for a many vortex state can be written \begin{equation}\label{5.30} \psi_{m}[z]=\prod_{p<q}(z_{p}-z_{q})^{\alpha+2m} \prod_{r} exp(-\mid z_{r} \mid^{2}/4l^{2}) \;\;, \end{equation} where $z_{p}$'s are positions of vortices. $m$ is a nonnegative integer so that the exponent $(\alpha+2m)$ provides a correct quantumstatistics. The density $n$ and the filling factor $\nu$ in such a state are \begin{equation}\label{5.40} \nu_{m}=2\pi l^{2} n_{m}=\frac{1}{\alpha+2m} \;\;. \end{equation} For $m=0$ the density is just $n_{0}=1/2\pi\alpha l^{2}=\rho_{s}/\alpha$. $\alpha$ is roughly equal to the number of electrons inside the core so in the $m=0$ state vortex cores would have to overlap slightly. Certainly this density is close to that of the Abrikosov lattice \cite{abr}. For larger $m$ the density is smaller and finally we should get outside of the crystalline regime. Then as the density of vortices is driven to change by the changes of the real magnetic field $B_{ext}$ we should observe some plateaux at the densities $n_{m}$ and maybe also at some other quantised filling factors. For conventional superconductors, which are mildly type $II$, $\alpha$ is large, so the difference between $n_{m}$ and $n_{m+1}$ is too small for the plateaux to be distinguishable experimentally. For extremaly type $II$ superconductors or high $T_{c}$ superconductors we can expect $\alpha$ to be even $\sim 1$. In such a case the first plateaux can be expected already at $\sim 1/3$ and $\sim 1/5$ of the Abrikosov lattice density. The main observation of this section is that if the Magnus force is translated into the language of interaction with some effective uniform magnetic field, the field appears to be even $\sim 10^{6}T/\AA$. It has nothing to do with the real magnetic field which is at best $\sim 10 T$. The FQHE is a result of this huge effective magnetic field and of repulsive intervortex potential but its existence does not depend on the vortex statistics as the FQHE is possible even for bosons \cite{jain}. However the pattern of the FQHE plateaux can help to identify the quantumstatistics. An experimental setup to detect the FQHE would consist of a planar sample of some pure superconducting material in external uniform perpendicular magnetic field. One would have to measure the total flux $\Phi$, which penetrates through the sample. This flux gives the actual number of vortices in the sample because each vortex carries one flux quantum. Just after continuously turning on weak magnetic field the field lines would be pushed out of the sample. Only above some threshold value of the flux the energy of the field could be minimised by creating a vortex line. The story would repeat until some stable FQHE plateau were reached. The plateau is a manifestation of a very stable many-vortex state so an addition of one vortex to this state would cost more energy then to a state far from the plateau. When the stable state is approached from low densities the addition of one more vortex should be much easier than usually. Quite opposite, when we lower the external magnetic field it is easier to remove one flux quantum from the sample just above the stable state and more difficult to remove it just below the state, see the figure. The moduli of the change in the driving flux $\Delta\Phi$ (or some equivalent quantity) necessary to change the flux through the sample by one quantum will develop a characteristic pattern around the stable density $n_{0}$. Two measurements, one with adiabatically increasing and one with adiabatically decreasing external magnetic field, would give two curves with opposite polarisation. Taking their difference will amplify the effects due to the plateau and remove the not neccesarily constant bias. It should be stressed here that our results are not in contradiction with predictions of a bosonic Hall effect in Josephson junction arrays \cite{jja}. The arrays are strictly planar devices. The penetration lenght $\Lambda$ is likely to be greater then the sample size. What is more, it seems to be possible to excite a vortex without any bound states in the core, which indeed might be a boson. Our results are exact in the limit of long straight-linear parallel vortices. In practice it is sufficient that the sample is thick enough for the penetration lenght $\Lambda$ to be close to that in the bulk. The sample thickness would have to be a bit more then $d=100\AA$. For such a $d$ there is still no space for vortex entanglement. In the absence of entanglement vortices can be uniquely projected on a plane. In nonzero temperature one should expect transversal modes to be excited. These excitations do not affect the Berry phase picked up during vortices exchange as it depends on the overall topological properties of vortices. Vortices excited to different transversal states although in principle distinguishable interact statistically. It is an example of mutual statistical interaction \cite{mut} introduced first to describe interlayer phase correlations in the double-layer Hall effect. The plateaux pattern for our anyonic vortices is given by (\ref{5.40}). In distinction to bosonic Hall effect the constant $\alpha$ is nonzero. \section{ Summary } We have given two new arguments why vortices in superconducting films should be anyons. In addition we have discussed two different experiments where this theoretical prediction could be verified. \paragraph{Aknowledgement.} I would like to thank Prof. B.Halperin for stimulating comments on Ref.\cite{anyon}. This research was supported in part by the KBN grant No. 2 P03B 085 08 and in part by Foundation for Polish Science fellowship.
\section{Introduction} Our best candidates, to date, of fundamental descriptions of nature, now all seem to be connected by the duality conjectures \cite{duff,schwarz,hull,witten}. The emerging picture is one of an underlying, even more fundamental, theory in which the other theories emerge as one approaches various limits in the moduli space. Much effort has gone into understanding the $N=4$ \cite{schwarz,hull}, $N=2$ \cite{sei,kach,klemm,louis} and more recently the $N=1$ \cite{seib,vaf,har} cases. In particular the $K3$ manifold plays a central role \cite{hull,witten,ferr,strom,klemm,louis,vaf,paul,town,sen} in many scenarios. Compactification of $d=11$ supergravity on a compact manifold of $G_{2}$ holonomy gives an effective four dimensional $N=1$ theory. Joyce has given many examples of such manifolds \cite{J1,J2}. It is natural to conjecture that these manifolds play a crucial role in the $N=1$ truncations of the conjectured dualities between the Type IIA superstring on $K3$, the heterotic string on $T^{4}$ and $d=11$ supergravity on $K3{\times}S^{1}$. In \cite{har} a manifold of $G_{2}$ holonomy was constructed by considering freely acting involutions of a $K3{\times}T^{3}$ orbifold and resolving the singularities. These involutions translate directly to actions on the heterotic side giving an example of $N=1$ duality. However, freely acting involutions of this type are very limited, and the case when the involutions act with fixed points on the seven coordinates on the supergravity side is less well understood. In fact, all of the manifolds of \cite{J1,J2} are constructed by resolving fixed point singularities, and if these manifolds are to play a role in $N=1$ duality then this situation needs to be understood. This work aims to shed some light on this problem. We work at a generic point in moduli space throughout the following. We focus on the simplest example given in \cite{J1}, which we denote by $J$, constructed with three non-freely acting involutions of the seven torus, the singularities of which are then resolved. In section two a brief outline of Joyce's construction is given. Then in section three we present the heterotic dual compactification. In particular, we are forced to consider a new kind of 'overlapping orbifold' on the heterotic side of the duality map. We end with some conclusions. \section{Joyce's Construction} In \cite{J1,J2} Joyce gave explicit constructions of 7-manifolds of $G_2$ holonomy. When eleven-dimensional supergravity is compactified on these manifolds, the resulting theory is four dimensional N=1 supergravity with $b_2$ vector multiplets and $b_3$ chiral multiplets, where $b_2$ and $b_3$ are the non-trivial Betti numbers of the 7-manifold. The examples of \cite{J1} are all constructed by orbifolding the seven-torus by various discrete isometries and then resolving the singularities by replacing them with non-compact Eguchi-Hanson geometries, a process that is now more than familiar to string theorists. The simplest example given in \cite{J1}, which we denote by $J$ is constructed as follows: Define the seven-torus coordinates as $(x_1,......,x_7)$. Three $Z_2$ isometries of $T^7$ are defined by: \begin{equation} \alpha(x_1,....x_7) = (-x_1,-x_2,-x_3,-x_4,x_5,x_6,x_7) \end{equation} \begin{equation} \beta(x_1,....x_7) = (-x_1,1/2-x_2,x_3,x_4,-x_5,-x_6,x_7) \end{equation} \begin{equation} \gamma(x_1,....x_7) = (1/2-x_1,x_2,1/2-x_3,x_4,-x_5,x_6,-x_7) \end{equation} Obviously each of these $Z_{2}$'s has 16 fixed singular $T^{3}$'s and each one defines an orbifold limit of a particular $K3{\times} T^{3}$. There are thus 48 fixed singular $T^{3}$'s of the surface. However one must ask how the other two isometries act on the singular set of each $Z_{2}$. In particular, the $Z_{2}{\times}Z_{2}$'s generated by $(\beta,\gamma)$, $(\alpha,\gamma) $ and $(\alpha,\beta)$ act freely on the singular sets of $\alpha$, $\beta$ and $\gamma$ respectively. This implies that each isometry has only four singularites of the original sixteen, giving a total of 12 singular $ T^{3}$'s. Resolving each of these is crucial for the consistency of the supergravity theory. This is done in the usual way for $K3$ singularities - by inserting Eguchi-Hanson geometries, \cite{eh}. The Betti numbers of the singular $T^{7}$ are $b_{2} = 0$ and $b_{3} = 7$. The resolution of each singularity adds 1 to $b_{2}$ and 3 to $b_{3}$. The Betti numbers of this Joyce manifold are thus $b_{2} = 12$ and $b_{3} = 43$. \section{The Heterotic Dual and Generalised \newline Orbifold} Eleven-dimensional Supergravity on $K3{\times}T^{3}$ is conjectured to be dual to the heterotic string on $T^{6}$. Orbifolding the $K3{\times}T^{3}$ by symmetries such as the Enriques involution on the $K3$ part translates to a particular orbifold action on the heterotic side. This has been illustrated successfully in several examples \cite{ferr,sen,vaf,har}. In particular, in \cite{har} an example was considered in which $T^{7}$ was modded out by a particular $Z_{2}{\times}Z_{2}{\times}Z_{2}$, one of which defined a particular orbifold limit of $K3{\times}T^{3}$. This is what has been called $\alpha$ in the construction of $J$. The remaining $Z_{2}$'s acted freely, and we note that they are essentially 'freely acting versions' of what we denoted by $\beta$ and $\gamma$ in constructing $J$. This is so because the example of \cite{har} contained half shifts of $S^{1}$'s defined by $x_{i} \rightarrow x_{i} + 1/2$. In that example there is only one $K3$ present in constructing the 7-manifold and it is straightforward to map the freely acting $Z_{2}{\times}Z_{2}$ to the heterotic side. So what is the dual of the supergravity theory on $ J$? $J$ was constructed by taking $T^{7}$, orbifolding by three $Z_{2}$'s ({\sl each of which defines a particular $K3\times T^{3}$ }) and resolving all the singularities. In a sense, we have three overlapping $K3$'s and when the singularities are suitably resolved we give non-trivial holonomy to the whole $T^{7}$, promoting it to $G_{2}$. This implies that on the heterotic side, the dual theory should be 'an overlapping' of three $Z_{2}{\times}Z_{2}$ orbifolds, because each of the three $K3$'s on the supergravity side should be treated on an equal footing. Denote by ${\alpha}\prime$, ${\beta}\prime$ and ${\gamma}\prime$ the action of the $Z_{2}{\times}Z_{2}$ generated by $(\beta,\gamma)$, $(\alpha,\gamma)$ and $(\alpha, \beta)$ respectively on the heterotic side. First we note that, if treated separately, the $Z_{2}{\times}Z_{2}$ orbifolds given by ${\alpha}\prime$, ${\beta}\prime$ and ${\gamma}\prime$ each produce the same massless spectra as the model condidered in \cite{har} \footnote {Appropriate $S^{1}$ shifts are included in the duality map, so that the adiabatic argument of \cite{vaf} applies to each orbifold}, namely four vector multiplets and 19 chiral multiplets. To find the massless spectrum of our 'overlapping orbifold', it is only necessary to note that we have three copies of the same entity on both sides of the duality map.{\sl On the supergravity side resolution of singularities of {\it each} $K3$ gave a specific massless spectrum. On the heterotic side, each of the three orbifolds involved in the 'overlapping' each give essentially the same compactification}. This suggests that we sum the massless spectra separately, which leads naturally to the definition of the 'overlapping orbifold' as one in which each orbifold involved in the process should be treated separately, and the spectra summed. However, we should only count the dilaton and six moduli once. This gives a spectrum with precisely 12 vector multiplets and 43 chiral multiplets. \section{Conclusions.} It is further interesting to note that all the examples of Joyce constructed from $Z_{2}$'s have the Betti numbers of the singular $T^{7}$ as $b_{2} = 0$ and $b_{3} = 7$. This should then always correspond to the seven chiral multiplets containing the dilaton and moduli. In our example, the number of singularities of each isometry then corresponded on the heterotic side to the number of surviving $ N=4$ vector multiplets, each of which give rise to one $ N=1$ vector multiplet and three $ N=1$ chiral multiplets on {\it both} sides. Of the more general cases considered in \cite{J2}, it turns out that for certain subsectors of the singularities there exists more than one topologically distinct ways of resolving. The different resolutions add different numbers to $b_{2}$ and $b_{3}$. This then should correspond on the heterotic side to subsets of $N=4$ vector multiplets surviving the orbifold projection and possibly to extra massless states from the twisted sectors. It is thus natural to conjecture that there is a heterotic dual for each of the manifolds given in \cite {J2} and this deserves further investigation \cite{ba}. \newline {\bf Acknowledgements}. The author is extremely indebted to Chris Hull, Wafic Sabra, and Tomas Ortin for discussions and in particular to Ashoke Sen for many illuminating conversations. The author would also like to thank PPARC, by whom this work is supported.
\section{Introduction} In recent years strongly interacting bosons attracted a lot of attention. Experiments on thin granular films show a zero temperature superconductor to insulator (SI) phase transition as a function of the film thickness~\cite{gold} or magnetic field~\cite{hpsi}, and experiments on Josephson junction arrays (JJA) exhibit the same SI transition as a function of the ratio between Josephson coupling and charging energy~\cite{jja}. These findings are interpreted in terms of a bosonic model~\cite{efe,doni,fwgf} in which the competition between the hopping of Cooper pairs and the Coulomb interaction between them is responsible for the SI transition. In relation to the effect of disorder on the SI transition a Bose-glass phase is discussed in Refs.~\cite{fwgf,f-db,zimrev}. An interesting extension of the minimal model with on-site interaction only is to account for finite-range interactions, which are present in Josephson junction arrays and $^4$He-films on substrates in two dimensions, and bulk $^4$He in three dimensions. Artificially fabricated JJA's are ideal model systems to study the superconductor to Mott-insulator transition as a function of the coupling constants or magnetic field~\cite{jja}. The finite range of the interaction in these systems leads to commensurability and frustration effects. Furthermore, coexistence phases may appear that are called supersolids. In these phases a charge-density wave (CDW) that is stabilized by the interaction may coexist with superfluidity. In other words the system has diagonal long-range order (LRO) and off-diagonal LRO at the same time. The supersolid phase was studied in the early seventies after Andreev and Lifshitz~\cite{andlif} suggested that vacancies in a quantum crystal such as $^4$He might Bose-Einstein condense at low temperatures without destroying the crystal structure, thereby establishing a superfluid solid~\cite{l-ss}, or supersolid. Normally bosons at zero temperature are either superfluid (with off-diagonal LRO) or solid (with diagonal LRO). However, for a finite range of the interactions between the bosons a coexistence phase was predicted within mean-field approximations~\cite{mt-ss,lf-ss,bfs,rs-ss}. Experiments have been performed on $^{4}$He, but no positive identification of this coexistence phase has yet been made. There are, however, hints towards such a phase \cite{m-ss,lg-ss}. Recently several other kinds of coexistence phases were studied. The possibility of a spontaneous vortex anti-vortex lattice in superfluid films was explored in Ref.~\cite{vavl} and a coexistence phase of superfluidity and hexatic orientational order was proposed in Ref.~\cite{hex}. Orientational order in incompressible quantum Hall fluids is discussed in Ref.~\cite{leon}. Collinear supersolids were studied in Refs.~\cite{gergy,rs-mc}. Finally, we mention the relation between 2D bosons and 3D flux-lines in type II superconductors (high $T_{c}$ materials) in a magnetic field~\cite{nelson,fl-d,feigel}. Also in these systems different kinds of LRO may coexist and the equivalent of the supersolid is discussed in Refs.~\cite{blatt,frey}. Related is the question whether or not vortices may form a disentangled liquid, which would imply a normal ground state for bosons with long-range Coulomb interaction~\cite{feigel}. In this paper both the SI transition and the supersolid phase will be studied in some detail. In the next section several related models for interacting bosons will be discussed, of which the Bose-Hubbard model is the most general. In the limit of infinite on-site interaction the Bose-Hubbard model reduces to the spin $\frac{1}{2}$ XXZ model. For a large number of bosons per lattice site the Quantum-Phase model is applicable. The phase diagram of the latter two models are first determined within mean-field theory in Section~\ref{sec-mf}. A variational Ansatz for the ground state wave function that treats all three models on equal footing~\cite{baltin} is presented in Section~\ref{sec-vm}. In Section \ref{sec-nm} more accurate results for the phase diagram will be obtained by means of Quantum Monte Carlo. Part of the data has already been published as Ref.~\cite{ow-ss}. Finally the paper closes with a discussion of the various results in Section~\ref{sec-d}. \section{Three model Hamiltonians} \label{sec-tmh} A convenient starting point for the description of interacting lattice bosons is the Bose-Hubbard Hamiltonian \begin{equation} H_{BH}=\frac{1}{2}\sum_{ij}n_{i}U_{ij}n_{j} -\mu\sum_{i}n_{i} -\frac{t}{2}\sum_{\langle ij\rangle}(b^{\dagger}_{i}b_{j}+ b^{\dagger}_{j}b_{i})\;, \label{bh} \end{equation} where $b^{\dagger}, b$ are the creation and annihilation operators for bosons that satisfy the commutation relation $[b_i,b_j^{\dagger}]=\delta_{ij}$ and $n_i=b^{\dagger}_ib_i$ is the number operator. The first term describes the density-density interaction between the bosons. We will take it to be short range, $U_{0}$ for on-site, $U_{1}$ for nearest neighbors, and $U_{2}$ for second-nearest neighbors. The second term describes the coupling of the density to a chemical potential, and the third describes the hopping from site to site with hopping integral $t$. The first two terms will cause the bosons to be localized if the interaction dominates over the hopping. In the opposite case where the hopping dominates the bosons will form a superfluid. In the limit of large on-site interaction $U_{0}$, only the states with 0 and 1 boson per lattice site survive, and by identifying $b^{\dagger} = S^{+} = S^{x}+iS^{y}$ and $n = S^{z}+\frac{1}{2}$ the mapping to the spin $\frac{1}{2}$ XXZ Hamiltonian is obtained. It reads \begin{equation} H_{XXZ}=\frac{1}{2}\sum_{i,j}S^{z}_{i}U_{ij}S^{z}_{j} -h\sum_{i}S^{z}_{i} -t\sum_{\langle ij\rangle}[S^{x}_{i}S^{x}_{j}+S^{y}_{i}S^{y}_{j}]\;, \label{xxz} \end{equation} where $h=\mu-\frac{1}{2}\sum_i U_{0i}$. This Hamiltonian has been investigated extensively on the mean-field level for nearest-neighbor as well as next nearest-neighbor interactions. In the limit of a large number of bosons per lattice site one may parameterize $b^{\dagger}=\sqrt{\rho}e^{i\varphi}$ and take $\rho$ equal to the boson density. Thus, only the phase remains as a dynamic variable and the Quantum-Phase Hamiltonian reads \begin{equation} H_{QP}=\frac{1}{2}\sum_{i,j}n_{i}U_{ij}n_{j} -\mu\sum_{i}n_{i} -J\sum_{\langle ij\rangle}\cos(\varphi_{i}-\varphi_{j})\;, \label{qp} \end{equation} where $J=\rho t$. The phase and number are canonical conjugate variables and satisfy $[\varphi_i,n_j]=i\delta_{ij}$. This is exactly the Hamiltonian for Josephson junction arrays with an applied gate voltage in order to tune the chemical potential \cite{bfkos}. It is convenient to introduce a parameter $n_0= \mu/\sum_{i}U_{ij}$, which allows us to rewrite the Coulomb interaction and chemical potential term in Eqs.~(\ref{bh}) and (\ref{qp}) as \begin{equation} H_{\mbox{\footnotesize int}}= \frac{1}{2}\sum_{i,j}(n_{i}-n_{0})U_{ij}(n_{j}-n_{0}) \;. \label{int} \end{equation} We expect that the properties of the Quantum-Phase model will be periodic in the variable $n_{0}$ with period 1. \section{Mean-field} \label{sec-mf} A first insight into the properties of these models may be gained by considering their mean-field solutions. The mean-field approximation for the spin model in Eq.~(\ref{xxz}) is straightforward and was carried out by various authors \cite{mt-ss,lf-ss,bfs}. The question of a supersolid phase in this model was first addressed in Ref.~\cite{lf-ss} and investigated further in Ref.~\cite{bfs}. We review the spin mean-field theory in Subsect.~\ref{subsec-smf}. A mean-field approximation for the Quantum-Phase model which is able to identify both diagonal LRO and off-diagonal LRO appeared in Ref.~\cite{rs-ss}. It is reviewed in Subsect.~\ref{subsec-qpm}. Mean-field approaches to the Bose-Hubbard Model have been discussed in Refs.~\cite{fwgf,skpr}, but to our present knowledge there exists no mean-field approximation for the Bose-Hubbard model which allows for the identification of a supersolid phase. In Sect.~\ref{sec-vm} we present a variational Ansatz for the many-body wave function that allows the three models to be treated on an equal footing. This is important, as it allows for a comparison between the models on the same level of approximation. Since we use a factorizable Ansatz for the wave function, it also has mean-field character. The validity of the mean-field approximation in two dimensions is often limited, since the lower critical dimension for models with continuous symmetries is two, which is the dimension we are particularly interested in. This is a serious problem at finite temperature where the models do not even exhibit superfluid long-range order as formulated in the theorem of Mermin and Wagner, and Hohenberg~\cite{rmwh}. Useful results may, however, be gained from mean-field approximations at zero temperature where quantum dynamics formally adds an extra dimension, the imaginary time axis. The systems then do have long-range order and mean-field results may be qualitatively correct. This will be shown in the present article by scaling arguments and confirmed by the Monte Carlo simulations. In the remainder of this section we present the different mean-field type approaches. \subsection{Spin Mean-Field} \label{subsec-smf} The spin mean-field theory can be formulated by linearizing the spin-spin interaction. This yields the standard self-consistency equations for the magnetization \begin{equation} \langle{\bf S}_{i}\rangle=-\frac{1}{2} \frac{{\bf H}_{i}}{\mid{\bf H}_{i}\mid} \tanh\left(\frac{\beta}{2}\mid{\bf H}_{i}\mid\right)\;, \end{equation} where the subscript $i$ refers to the sublattice and $\beta$ is the inverse temperature (we use units in which $k_{B}=\hbar=c=1$). In order to identify spatially varying solutions one needs to introduce several sublattices, at least three for interactions including nearest and next-nearest neighbors. The symmetry in the $xy$-plane reduces the number of independent variables and most of the phase boundaries can be determined analytically. If we allow for three different sublattices ($A$, $B$ and $C$), the effective field is given by \begin{equation} {\bf H}_{A}=(-4t\langle S^{x}_{B}\rangle, -4t\langle S^{y}_{B}\rangle,2U_{1}\langle S^{z}_{B}\rangle +2U_{2}\langle S^{z}_{C}\rangle-h) \;, \end{equation} and corresponding equations for ${\bf H}_{B}$ and ${\bf H}_{C}$. The connection to hard-core bosons is made by remembering that $S^z$ corresponds to the particle number, and $S^{x/y}$ to the boson creation operators. A staggered magnetization in $z$-direction implies a solid ( charge-density wave) ordering of the bosons. Finite magnetization in $xy$-direction translates into off-diagonal LRO for the boson model. The coexistence phase of both types of LRO is the supersolid phase for the bosons. The result is shown in Fig.~\ref{spinfig} for $U_{2}=0.1\;U_{1}$. Two different types of solid ordering occur for different chemical potential, i.e., $(\pi,\pi)$ ordering around half-filling, and additional $(\pi,0)$ and $(0,\pi)$ ordering around quarter filling. \subsection{Quantum-Phase Model} \label{subsec-qpm} Now we turn to the mean-field treatment of soft-core bosons following Refs.~\cite{rs-ss,khw-d}. After decoupling the hopping and the interactions between the different sites in Eq.~(\ref{qp}), one arrives at a local problem \begin{equation} H_{mf}^i\;|\psi_i\rangle = E_i\;|\psi_i\rangle \; , \end{equation} with the $2\pi$-periodic solutions for the wave function in the phase representation \begin{equation} \psi_i(\varphi_i) = \exp(i\eta_i x_i) \; f_i(x_i) \; , \label{eq:sol} \end{equation} where $\varphi_i=2\,x_i$, $\eta_i= 2\,n_{0} \;-\; \frac{2}{U_0} \sum_{j \neq i} U_{ij}(\langle n_{j}\rangle-n_{0})$. After these substitutions the following Mathieu equation for $f$ is obtained \begin{equation} \left\{ \frac{\partial^2}{\partial x_i^2} \; +\;e_i\;+\;2\,r_i\cos(2\,x_i) \right\} f_i(x_i) \;=\;0 \end{equation} \begin{equation} \mbox{with}\ \ \ r_i\;=\;\frac{4J}{U_0} \sum_{\beta}\langle\cos(\varphi_{i+\beta})\rangle \; . \end{equation} The sum runs over nearest neighbors $\beta$ of $i$. The parameter $e_i$ is proportional to the energy and should be minimized. The ground state is defined via the self-consistency equation \begin{eqnarray}\nonumber \langle n_i\rangle\ \ \ \ &=&\ \ \ \langle\psi_i^0\,|\;n_i\; |\psi_i^0\rangle\\ \langle\cos(\varphi_i)\rangle&=&\langle\psi_i^0\,|\, \cos(\varphi_i)\,|\psi_i^0\rangle \;. \label{sc} \end{eqnarray} In addition the free energy has to be minimized in order to have an unambiguous solution of Eq.~(\ref{sc}). The numerical solution of this set of equations is straightforward. In order to allow for spatially varying solutions we again define sublattices. Our numerical solutions confirm the results of Ref.~\cite{rs-ss}, however, some details differ. For on-site interaction we obtain Mott insulating lobes centered around integer values $N$ of $n_0$, see Fig.~\ref{mottfig}. In these incompressible lobes the density is pinned to $N$. At half integer values of $n_0$ the critical value of the hopping vanishes, due to the degeneracy in energy of states with $N$ and $N+1$ particles per site. Including nearest-neighbor interaction the calculation has to be performed on two sublattices. We obtain two different kinds of insulating lobes, one with integer filling and homogeneous density, another with half-integer filling and a checkerboard charge-density wave, or solid order, centered around half integer values of $n_0$. In addition we find a region of supersolid phase around the half lobe where the checkerboard charge-density wave and superfluidity coexist, see Fig.~\ref{cbssfig}. Again, the superfluid phase occupies the parts on the right of the phase diagram and also between the insulating lobes down to zero hopping, in slight contrast to the result of Ref.~\cite{rs-ss}. The fraction of supersolid phase in the phase diagram depends on the ratio $U_{1}/U_{0}$, see Fig.~\ref{numbfig}. In the hard-core limit $U_0 \rightarrow \infty$ the supersolid phase vanishes. An expansion of the ground state wave function $|\psi_i^0\rangle=\sum_{n}c^{i}_{n}|n\rangle$ into particle number eigenstates $|n\rangle$ yields more information about the nature of the supersolid phase. The inset of Fig.~\ref{numbfig} shows $|c_{2}|^{2}$ at $n_{0}$=0.5 as a function of $J/U_{0}$. The figure indicates that double occupancy is important for the supersolid phase. Thus, we conclude that the supersolid phase is favored by fluctuations in the particle number. The inclusion of further interactions leads to more structure in the phase diagram. For next-nearest neighbor interaction we introduce three sublattices. As a result, The phase diagram shows a lobe with quarter filling and a phase where this quarter structure and superconductivity coexist, see Fig.~\ref{nnnfig}. \subsection{Hard-core vs. Soft-core} \label{subsec-hcsc} The previous subsections showed two extreme cases of the Bose-Hubbard model, the hard-core limit and the limit of large particle numbers. Although the phase diagrams look similar, these two limits behave qualitatively different as far as the supersolid phase is concerned. In the hard-core limit the existence of the supersolid is related to a finite next-nearest neighbor interaction. The supersolid does not exist in the model including nearest-neighbor interaction only~\cite{bfs}. Furthermore, there is no supersolid phase at exactly half filling. This leads to the following interpretation of the supersolid phase. At densities corresponding to half filling the particles form an incompressible solid. Away from half filling superfluidity is enabled by defects in the solid structure that Bose-Einstein condense. This interpretation does, however, not explain why next-nearest neighbor interaction is necessary for the formation of the supersolid. In the limit of many particles per site the Quantum-Phase model applies. In this case the supersolid phase already exists for nearest-neighbor interaction only and at half filling. This is related to excitations which are forbidden in the hard-core limit. We observe that a large nearest-neighbor interaction $U_{1}$ or small on-site interaction $U_{0}$ favors the supersolid, in the hard-core limit the supersolid is suppressed. In the soft-core limit the supersolid phase is also present at half-filling at the tip of the checkerboard lobe. Thus, it seems that the system itself generates the defects (particle-hole excitations) that Bose condense, thereby turning the solid into the supersolid. \section{Variational Method} \label{sec-vm} Here we present a method which allows us to treat the three models within one scheme~\cite{baltin}. The idea is to guess a variational wave function for the problem \cite{gutz}. We use a variational wave function that is inspired by earlier successful Gutzwiller-like treatments of the spin model~\cite{huse} and the Bose-Hubbard model~\cite{roko,kraut}. The occupation number representation is well suited for all three models. Explicitly we resolve the identity as \begin{equation} 1=\sum_{\{n_i\}}|\{n_i\}\rangle\langle\{n_i\}| \end{equation} and express the trace as \begin{equation} \mbox{Tr ...}=\sum_{\{n_i\}}\langle\{n_i\}| ... |\{n_i\}\rangle, \end{equation} where the $n_i$ are in $[0,1]$ for the spin $\frac{1}{2}$ XXZ model, in $[0..\infty]$ for the Bose-Hubbard (BH) model, and in $[-\infty..\infty]$ for the Quantum-Phase (QP) model. Besides the difference in the allowed values of $n_i$, the matrix elements differ for the three models, e.g., \begin{eqnarray}\nonumber \mbox{BH model:}\;\;\;\;\;\; & & \langle\{n_i'\}|b_k|\{n_i\}\rangle = \langle\{n_{i\neq k}'\}|\{n_{i\neq k}\rangle \langle n_k'|n_k-1\rangle \sqrt{n_k} \\ \mbox{QP model:}\;\;\;\;\;\; & & \langle\{n_i'\}|\exp{i\varphi_k}|\{n_i\}\rangle = \langle\{n_{i\neq k}'\}|\{n_{i\neq k}\rangle \langle n_k'|n_k-1\rangle \end{eqnarray} As our variational wave function we take the product of single-site wave functions. This amounts to neglecting certain correlations and is therefore similar to a mean-field approximation. We use the following Ansatz $$ |\psi\rangle=\prod_{i=1}^{N}\frac{1}{\sqrt{Z_{i}}} \sum_{\{n_{i}\}}\exp\left\{-k_{i} (n_{i}-m_{i})^{2}/2\right\}|n_{i}\rangle $$ \begin{equation} \mbox{where }\ \ \ \ Z_{i}= \sum_{\{n_{i}\}}\exp\left\{-k_{i}(n_{i}-m_{i})^{2}\right\} \;, \label{ansatz} \end{equation} and $k_{i}$ and $m_{i}$ are real variational parameters. The single-site wave function is a superposition of boson number eigenstates weighted with a Gaussian. The average is controlled by $m_i$, the width by $k_i$, and spatial variations (different sublattices $i$) are allowed. By minimizing the energy expectation value $E=\langle\psi|H|\psi\rangle$ with respect to $k_{i}$ and $m_{i}$ we obtain the ground state wave function within our approximation. In the following, expectation values are understood to be calculated with the ground state wave function. As the discrete sums can in general not be performed analytically, we treat the problem numerically. For on-site and nearest-neighbor interactions two sublattices arranged in a checkerboard configuration are introduced. We define the superconducting order parameter as $\langle\exp(i\varphi)\rangle$ for the QP model, $\langle b\rangle$ for the BH model, and $\langle S^{-} \rangle$ for the XXZ model. The structure factor \begin{eqnarray}\label{eqstr} S(\vec{q}) & = & \bigg(\frac{1}{N}\bigg)^{2}\sum_{i,j} \exp[i\vec{q}\cdot(\vec{r}_{i}-\vec{r}_{j})]\langle n_{i}n_{j}\rangle \end{eqnarray} yields information about the solid order. A non-vanishing structure factor at finite $\vec{q}$ signals diagonal LRO. By decomposing the lattice into two sublattices we gain information about $S(\pi,\pi)$. A finite $S(\pi,\pi)$ corresponds to a checkerboard arrangement of the particles. For next-nearest neighbor interactions we introduce four sublattices which yields additional information about $S(\pi,0)$ and $S(0,\pi)$. Figure~\ref{qpvarfig} shows the phase diagram for the QP model with on-site and nearest-neighbor interaction obtained with our variational Ansatz. Both the superfluid-insulator and the crystalline order transition can be identified and agree with previous mean-field calculations \cite{rs-ss,ow-ss}. The phase boundaries are periodic in $n_{0}$. We find that superfluidity sets in simultaneously on both sublattices $A$ and $B$. Within the supersolid phase the value of the superconducting order parameter on $A$ differs from that on $B$, but both are nonzero. Beyond the crystalline order phase boundary the order parameter does not exhibit any difference on the sublattices. For the XXZ model the variational method confirms that there is no supersolid for on-site and nearest-neighbor interaction only. In the presence of next-nearest neighbor interaction we find supersolid phases in perfect agreement with \cite{bfs} and Fig.~\ref{spinfig}. The phase diagram for the BH Hamiltonian is shown in Fig.~\ref{bhvarfig}. The size of the lobes decreases with increasing $n_0$. At point $\alpha$ the supersolid vanishes. This might be due to the lower bound for the occupation numbers $n\geq 0$. For small $n_0$ charge fluctuations are suppressed. Hence, we conclude that charge fluctuations are necessary for the supersolid. For large $n_{0}$ the phase boundaries of the QP model approach those of the BH model, if the hopping is rescaled according to Section~\ref{sec-tmh}, i.e., if $t$ and $J/n_{0}$ are identified. \section{Quantum Monte Carlo} \label{sec-nm} Several methods are available for determining the phase diagram for the three model Hamiltonians by numerical means. Exact diagonalization is possible for the XXZ model and small systems, i.e., less than about 26 lattice sites. Furthermore it is possible to perform Monte Carlo simulations for all the three models on much larger systems. The Bose-Hubbard model was studied in one dimension in Refs.~\cite{bsz,nfsb}, and one data point for the phase boundary was obtained in two dimensions in Ref.~\cite{kt-mc}. The Bose-Hubbard model was more recently studied in the context of a supersolid phase in Ref.~\cite{gergy}. The two-dimensional Quantum-Phase model was studied in Ref.~\cite{swgy} in relation to the universal conductivity at the SI transition and in Refs.~\cite{rs-mc,ow-ss} in relation to supersolid phases. In this section we present more results on Monte Carlo simulations of the two-dimensional QP model in addition to those in Ref.~\cite{ow-ss}. We discuss the on-site problem and compare to the analytic results of $t/U$ expansions \cite{fm-bh,fm-ex}. The inclusion of nearest-neighbor interactions allows us to study the supersolid phase. \subsection{Duality Transformation} \label{subsec-dt} In order to study the Quantum-Phase model by means of Monte Carlo we map the 2D quantum model onto a (2+1)D classical model of discrete divergence-free currents. The essential feature of this mapping were presented in Ref.~\cite{fl-d} and the derivation makes use of duality transformations that were developed in Ref.~\cite{jkkn}. We start from the basic expression for the partition function $Z$ for the Quantum-Phase model of Eq.~(\ref{qp}) \begin{equation} Z=\mbox{Tr}\;\exp(-\beta H_{QP})\;, \label{az} \end{equation} where $\beta$ is the inverse temperature. We go over to a Euclidean path-integral formulation by introducing time-slices, i.e., dividing $\beta$ in $N_{\tau}$ intervals of size $\epsilon$, such that $N_{\tau}\epsilon=\beta$. Inserting complete sets of states at each time slice we arrive at $$ Z=\sum_{\{n_{i,\tau}=0,\pm1,\pm2,\cdots\}} \int{\cal D}\varphi_{i,\tau}\exp{\Big\{ -\frac{\epsilon}{2}}\sum_{ij,\tau} n_{i,\tau}U_{ij}n_{j,\tau}+ \epsilon\mu\sum_{i,\tau}n_{i,\tau} $$ \begin{equation} +i\sum_{i,\tau}n_{i,\tau}\dot{\varphi}_{i,\tau}+ \epsilon J\sum_{\langle ij\rangle,\tau} \cos(\varphi_{i,\tau}-\varphi_{j,\tau}-A_{ij}){\Big\}}\;. \label{laz1} \end{equation} We included a coupling to a vector potential $A_{ij}=(2\pi/\Phi_{0}) \int^{j}_{i}{\bf A}\cdot d{\bf l}$ for later convenience. In order to proceed, we make one approximation, the Villain approximation~\cite{vill} for the cosine term. This amounts to expanding the cosine around all its minima \begin{equation} \exp\{\epsilon J\cos(\varphi_{i\tau}-\varphi_{j\tau}-A_{ij})\} \approx\!\!\!\sum_{\{m_{ij,\tau}\}} \exp\{-\frac{\epsilon J F(\epsilon J)} {2} (\varphi_{i\tau}-\varphi_{j\tau}-2\pi m_{ij,\tau}-A_{ij} )^{2}\}\;, \label{eq:av} \end{equation} where $m$ is a directed discrete field that lives on the bonds between lattice sites. The function $F$ is determined by demanding that the first two Fourier coefficients of the expressions on the left and right hand side of Eq. (\ref{eq:av}) are equal~\cite{vill} and is given by \begin{equation} F(x)=\frac{1}{2x\log\{I_{0}(x)/I_{1}(x)\}}\;. \label{vill} \end{equation} For large arguments $F$ approaches 1, as is clear from a direct expansion of the cosine potential. After a subsequent Poisson resummation, i.e., writing $$ \sum_{\{m_{ij,\tau}\}}f[m_{ij,\tau}]= \sum_{\{{\cal J}_{ij,\tau}\}}\int {\cal D}m f[m_{ij,\tau}] \exp(2\pi i\sum_{ij,\tau}m_{ij,\tau}{\cal J}_{ij,\tau})\;, $$ an integration over the fields $m_{ij,\tau}$ and the phases $\varphi_{i,\tau}$ yields a representation in terms of divergence-free discrete current loops $$ Z=\left[\epsilon JF(\epsilon J)\right]^{-1} \sum_{\{{\cal J}^{\nu}_{i,\tau}\}} \delta(\nabla_{\nu}{\cal J}^{\nu})\exp\{-S[{\cal J}]\}\;, $$ $$ S[{\cal J}]=\frac{\epsilon}{2}\sum_{ij,\tau} {\cal J}^{\tau}_{i,\tau} U_{ij} {\cal J}^{\tau}_{j,\tau} -\epsilon\mu\sum_{i,\tau}{\cal J}^{\tau}_{i,\tau} +\frac{1}{2\epsilon J F(\epsilon J)}\sum_{i,\tau,a=x,y} \left({\cal J}^{a}_{i,\tau}\right)^{2} -i\sum_{i,\tau}\vec{A}_{i}\cdot\vec{{\cal J}}_{i,\tau} $$ \begin{equation} =\sqrt{\frac{2}{K}}\left\{\sum_{ij\tau}({\cal J}^{\tau}_{i,\tau}-n_{0}) \left(\delta_{ij}+\frac{U_{1}}{U_{0}}\delta_{\langle ij\rangle}\right) ({\cal J}^{\tau}_{j,\tau}-n_{0})+\!\!\sum_{i\tau,a=x,y}\!\! \left({\cal J}^{a}_{i,\tau}\right)^{2}\right\} -i\sum_{i,\tau}\vec{A_i}\cdot\vec{{\cal J}}_{i,\tau}\;, \label{al} \end{equation} where $\nu=x,y,\tau$. The effective coupling constant is given by $\sqrt{2/K}$ with $K=8F(\epsilon J)J/U_{0}$, where $\epsilon$ is defined implicitely by $\epsilon=1/\sqrt{U_0F(\epsilon J) J}$. The time components ${\cal J}^{\tau}_{i,\tau}$ of the current correspond to the boson numbers $n_{i,\tau}$ along their world lines. We keep track of the energy dependence of the determinant that arises from integrating out the $m_{ij,\tau}$. This is relevant for determining the energy density and specific heat by means of Monte Carlo. For zero magnetic field ($\vec{A}=0$) the action in the representation of Eq.~(\ref{al}) is real, and therefore suitable for Monte Carlo methods. Using the standard Metropolis algorithm we generate configurations of currents in a system of size $L\times L\times L_{\tau}$. The condition that the currents ${\cal J}^{\nu}$ be divergence-free is taken into account by making Monte Carlo steps that preserve the property $\nabla_{\nu}{\cal J}^{\nu}=0$. Thus, we create or annihilate small current loops at every lattice site in all three directions, as well as periodic current loops that go through the whole system which is taken to have periodic boundary conditions in all three directions. The generation of configurations may be done in a canonical as well as in a grand-canonical way. Here we work in the grand-canonical ensemble in order to make contact to the mean-field phase diagrams. In the last line of Eq.~(\ref{al}) we used $\epsilon=1/\sqrt{U_0F(\epsilon J) J}$. With this choice the couplings are isotropic for on-site interaction, leading to an efficient numerical algorithm. The choice of the time lattice spacing $\epsilon$ needs some justification. Introducing time slices in the partition function Eq.~(\ref{laz1}) is exact in the limit $\epsilon \rightarrow 0$. This would produce extremely anisotropic couplings in the current-loop model and the numerical simulation would become impossible. Thus, a more reasonable choice for the time lattice spacing is necessary. The plasma frequency $\omega_{p}=\sqrt{JU_0}$ is a natural frequency for spin waves. A cutoff beyond this frequency by introducing a time spacing $\epsilon \approx 1/\omega_{p}$ should not do any harm. We tested the dependence on the choice of $\epsilon$ for the quantities of interest and observed that they are not sensitive to variations of $\epsilon$ by one order of magnitude, provided that the temperature is kept constant, i.e., $\beta=1/T=const.=N_{\tau}\epsilon$. Additional justification for the choice of a finite time spacing $\epsilon$ in the present paper is given by the fact that we are only interested in the critical regime, where a high-frequency cutoff should be irrelevant. \subsection{Finite-Size Scaling} The mapping to the current-loop model Eq.~(\ref{al}) is well suited for simulating the behavior of the Quantum-Phase model with short range interactions. In practice we are able to study system sizes up to $ 12 \times 12 \times 12$, or $10 \times 10 \times 25$, respectively. The factor that limits the largest systems to have $L$= 12 is the necessity to make the periodic current loops in the spatial directions. Only they can change the superfluid stiffness, which is a topological quantity that cannot be changed by making local current loops. As phase transitions take place only in the thermodynamic limit, we need an additional ingredient that relates the data obtained on finite-size systems to infinite system size properties as the critical coupling and exponents. This is provided by finite-size scaling. Let us first consider the onset of superfluidity in the system. Therefore we study the behavior of the superfluid stiffness $\rho_0$, a quantity which measures the response to a twist in boundary conditions for the phase. A finite value of $\rho_0$ in the thermodynamic limit reflects long-range phase coherence, i.e., off-diagonal LRO or superfluidity. A twist in the boundary conditions for the phases by an angle $\Theta$ increases the free energy density by \begin{equation} \Delta f=\frac{1}{2}\rho_0\left(\frac{\Theta}{L}\right)^2 \label{df} \end{equation} The twist may be realized by applying a magnetic field. In general the frequency and wave-vector dependent stiffness $\rho(k,\omega_{\mu})$ can be defined as \begin{equation} \rho(k,\omega_{\mu})=\left(\frac{1}{2e}\right)^{2} \sum_{i,\tau}\frac{-\delta^{2}\ln Z} {\delta A^{x}_{i,\tau}\delta A^{x}_{0}} \Bigg|_{\vec{A}=0} e^{-i\omega_{\nu}\tau-ikr_{i}}\;. \end{equation} The zero-frequency and zero wave-vector component of $\rho(k,\omega_{\mu})$ defines the superfluid stiffness. In terms of the currents it reads \begin{equation} \rho_0=\rho_{(k=0,\omega_{\mu}=0)}=\frac{1}{L^d L_{\tau}} \left\langle \left(\sum_{i,\tau}{\cal J}^x_{i,\tau}\right)^2\right \rangle=\frac{1}{L^{d-2} L_{\tau}}\langle w^2_x \rangle\,. \end{equation} The winding numbers $w_x=\frac{1}{L} \sum_{i,\tau} {\cal J}^x_{i,\tau}$ can be measured easily; they are only modified by periodic current loops, local loops do not change their value. The symmetry in $x/y$-direction further simplifies the numerics. Direct measurement in the critical regime is ruled out by the divergence of the correlation length $\xi$ at the continuous transition. The powerful method of finite-size scaling, however, takes advantage of this fact and allows us to study the critical behavior. Near a continuous phase transition the diverging coherence length is cut off by the finite system size $L$. As a result all quantities will depend on the ratio $\xi/L$. A finite-size scaling relation for the superfluid stiffness $\rho_0$ is readily derived~\cite{cha}. Assuming hyperscaling, the critical part of the free energy density behaves as the inverse correlation volume $\xi^d\xi_{\tau}$. Here, $\xi$ is the correlation length in real space which may be different from the correlation length $\xi_{\tau}$ in imaginary time, depending on the quantum dynamics of the model. The dynamical critical exponent $z$ can be introduced through $\xi_{\tau}\propto\xi^z$. Combining Eq.~(\ref{df}), the hyperscaling relation and the fact that $\xi$ is cut off by $L$ ($\xi_{\tau}$ by $L_{\tau}$) we arrive at the scaling behavior \begin{equation} \rho_{0}={L^{2-d-z}}\tilde{\rho} (b L^{1/\nu}\delta, L_{\tau}/L^{z})\,. \label{rhosc} \end{equation} Here, $b$ is a non-universal scale factor and $\tilde{\rho}$ is a universal scaling function with a smooth dependence on $L_{\tau}/L^{z}$ and $\xi/L=\delta^{-\nu}/L$. We assumed a power law critical behavior $\xi \propto \delta^{-\nu}$ with dimensionless distance to the transition $\delta=(K-K^{\ast})/K^{\ast}$ and critical coupling $K^{\ast}$. At the critical point ($\delta=0$), $L^{d+z-2}\rho_{0}$ is a function of $L_{\tau}/L^{z}$ only. Thus, plots of $L^{d+z-2}\rho_{0}$ vs.\ $K$ will intersect at the transition if $L_{\tau}/L^{z}$ is kept constant, with no further fitting involved. Furthermore, the data for $L^{d+z-2}\rho_{0}$ plotted as a function of $L^{1/\nu}\delta$ for different system sizes should collapse onto one single curve. This allows the exponent $\nu$ to be determined. As we are interested also in diagonal LRO and a possible supersolid phase, we measure the structure factor as well. The structure factor is the $(\pi,\pi)$-component of the equal time density-density correlation, see Eq.~(\ref{eqstr}), and may be derived as the second variation of the free energy with respect to a staggered chemical potential. Expressed in terms of the currents ${\cal J}^{\tau}$ it reads \begin{equation} S_{\pi}=\frac{1}{L^{4}L_{\tau}}\Big\langle\sum_{ij,\tau}(-1)^{i+j} {\cal J}^{\tau}_{i,\tau}{\cal J}^{\tau}_{j,\tau}\Big\rangle\;. \end{equation} The scaling relation for the structure factor is \begin{equation} S_{\pi}=L^{-2\beta/\nu} \tilde{S}(b'L^{1/\nu}\delta,L_{\tau}/L^{z})\;. \end{equation} This scaling relation arises, since the structure factor $S_{\pi}$ is related to the square of the staggered magnetization order parameter $M=\sum_{i}(-1)^{i}{\cal J}^{\tau}_{i}$ for which the exponent $\beta$ is the order parameter exponent. Near the phase transition, $M\sim \delta^{\beta}$ and $\xi\sim \delta^{-\nu}$, from which the quoted form follows. From a three-parameter fit to the scaling relation for different system sizes, the exponents $\nu$ and $\beta$ as well as the critical coupling constant $K^{\ast}$ are determined. Again, plots of $L^{2\beta/\nu}S_{\pi}$ will cross at the critical point if the ratio $L_{\tau}/L^{z}$ is kept constant. In both scaling relations a knowledge of $z$ is assumed. This is well-known for the SI transition and the scaling of $\rho_0$, i.e., in general $z$=2, except for the tips of the lobes where $z$=1 as dictated by particle hole-symmetry. This argument is based on a coarse-graining treatment \cite{doni} that allows for the explicit construction of a Ginzburg-Landau-Wilson free energy functional for the superconducting order parameter \cite{fwgf,owfs,jk-cg,kz-cg}, from which $z$ can be read off. For the transition related to the structure factor the situation is less clear. We were, for instance, not yet successful in explicitly constructing the Ginzburg-Landau-Wilson free energy functional for the CDW transition. From symmetry arguments one expects $z$=1 at the mean-field level~\cite{bafr}. Also the spin-wave analysis of Ref.~\cite{gergy} predicts $z$=1. However, it is unclear whether or not $z$ for the CDW will be renormalized by the coupling to the gapless sound mode in the superfluid part. As a working hypothesis we take $z$=1 and comment later on the consistency of this assumption. \subsection{Numerical Results} \label{subsec-nm} We report about the SI transitions for on-site interaction and nearest-neighbor interaction in the complete range of the chemical potential. A phase diagram is mapped out and the scaling predictions are verified. The existence of a supersolid phase in a finite region of the phase diagram is firmly established. The case of {\it on-site interaction} has been extensively studied in the past years. A phase diagram consisting of lobe-like structures in the $t-\mu$ or $J-n_{0}$ plane is found. Mott-insulating lobes are centered around integer values of $n_0$. This has been observed in mean-field approximations, $t/U$ expansions \cite{fm-bh,fm-ex}, and in Monte Carlo simulations of the one-dimensional Bose Hubbard model \cite{bsz}. The shape of the lobes in two dimensions is yet unknown. Therefore we first map out this phase diagram which is shown in Fig.~\ref{mottfig}. We performed the simulations for fixed $n_0$ and tuned through the transition by varying the effective coupling $K$. In the phase diagram in Fig.\ \ref{mottfig} this corresponds to moving on horizontal lines through the phase transition. At integer values of $n_0$ the system is particle-hole symmetric and we have $z=1$. According to the scaling form of Eq.~(\ref{rhosc}) we use $L=L_{\tau}$ which keeps the second argument of the scaling function constant. Four different system sizes with $L=6,8,10,12$ where studied. For the largest systems we typically use $10^6$ sweeps through the system for equilibration and measurement. In each sweep we try all possibilities of local and periodic current loops. Without any fitting the curves of $\rho_0 L^{d+z-2}=\rho_0 L$ vs.\ $K$ should cross in the critical point $K^*$. This serves as a good test of the scaling prediction. The critical exponent $\nu$ can be extracted by fitting the data of different system sizes near the transition to one scaling curve. This is shown in Fig.~\ref{dat1fig} where we plot the data as a function of $\delta L^{1/\nu}$. For $n_0=0$ we obtain the critical coupling $K^*=0.886\pm0.003$, and using a $\chi^2$ fit $\nu=0.69\pm0.06$. The exponent $\nu$ agrees with the expected 3D-XY behavior, where $\nu\approx 2/3$. Away from the symmetry points at the tips of the lobes the dynamical critical exponent is $z=2$. In order to keep $L_{\tau}/L^z$ constant we simulate systems with $L_{\tau}=L^2/4$, i.e., $L\times L \times L_{\tau}=6 \times 6 \times 9$, $8 \times 8 \times 16$ and $ 10\times 10\times 25$. Now the curves of $\rho_0 L^2$ should cross in the critical point. We obtained data for $n_0=$0.05, 0.1, ..., 0.45 as shown in Table \ref{table-ons}. From a fit to the scaling form we find $0.42<\nu<0.52$ with errors of $\approx 20 \%$. This is in agreement with the expected mean-field transition where $\nu=1/2$. A more accurate determination of the critical exponents turns out to require an enormous numerical effort with our method and is beyond the scope of our present work. In Fig.~\ref{mottfig}, the Monte Carlo data are shown together with mean-field phase boundaries and the results of the $t/U$-expansion of Refs.~\cite{fm-bh,fm-ex} in the limit of large $n_{0}$ (or $\mu$). A lobe shape is observed, these lobes are sharper than predicted by mean-field, but smoother than expected from the $t/U$-expansion. Approaching $n_0=0.5$ we observe that the critical value of $J/U_0$ decreases and a simple extrapolation yields $J^{\ast}=0$ for $n_0=0.5$. A numerical answer to the question whether $J^{\ast}$ is zero at $n_0=0.5$ is not possible, since the acceptance rates are very small for small $J/U_0$. The inclusion of {\it nearest-neighbor interactions} yields a richer phase diagram. The scaling for the superfluid stiffness is not modified by the inclusion of short-range interactions, i.e., the universality class is preserved. We expect $z=1$ in the case of particle-hole symmetry, $z=2$ otherwise. For nearest-neighbor interaction the lines with integer and half-integer values of $n_0$ exhibit particle-hole symmetry. For the transition related to charge-density wave order, we expect $z=1$ throughout the phase diagram. Guided by the mean-field phase diagram we expect the existence of two different kinds of lobes, with homogeneous particle densities centered around integer values of $n_0$, and with a checkerboard charge-density wave centered around half integer values of $n_0$. The supersolid phase is expected in the vicinity of the checkerboard lobe. For $n_{0}$= 0 and 0.5 we simulated $L\times L\times L$ systems, where $L$= 6, 8, 10, 12, as suggested by particle-hole symmetry and $z=1$. For $n_{0}$=0.1, 0.2 and 0.4 we obtain $z=2$ for the superfluid transition. In order to keep the ratio $L_{\tau}/L^{z}$ constant, we simulated $L\times L\times L^{2}/4$ systems, where $L$= 6, 8, 10. For $n_{0}$=0.4 near the transition for the diagonal LRO we performed the calculations with both $z=1$ and $z=2$. The results are summarized in Fig.~\ref{dat2fig} and Fig.~\ref{fig:zdisc} and Table \ref{table-nn}. In the following we discuss our data in detail. First we present our data for $n_{0}$=0.5. Figure \ref{dat2fig} shows that there are two separate transitions for diagonal and off-diagonal LRO with a coexistence region in between where {\it both} the superfluid density {\it and} the structure factor scale to a finite value in the thermodynamic limit. This demonstrates the coexistence of diagonal LRO and off diagonal LRO for soft-core bosons with nearest-neighbor interaction in 2 dimensions. Table~\ref{table-nn} shows that the exponent $\nu$ is different for the two transitions. For the transition related to superfluidity (point $\beta$ in Fig.~\ref{cbssfig}) we find $\nu=0.65\pm0.08$ which is consistent with the 3D XY universality class. For the transition related to crystalline order (point $\gamma$) the universality class is not known. We find $\nu\approx$0.55 and $\beta\approx$0.21. Also at $n_{0}$=0.4 we find two separate transitions that are the boundaries for the supersolid phase in between, see Table \ref{table-nn}. As compared to $n_{0}$=0.5 both transitions are shifted to smaller values of the coupling constant $K$. This is consistent with the mean-field phase diagram. Again the two transitions have different critical exponents. The transition related to superfluidity (the line separating phases ``Sol'' and ``SSol'' in Fig.~\ref{cbssfig}) has $\nu=0.44\pm0.08$ which is consistent with a mean-field transition in $d+z=4$ effective dimensions. For the transition related to crystalline order (between phases ``SSol'' and ``SF'') we compare the data for $z=2$ and $z=1$ in Fig.~\ref{fig:zdisc}. A fit to the expected scaling behavior is equally possible in both cases. The statistical errors of the data for $z=2$ are larger, since the simulation for $z=2$ requires larger $L_{\tau}$ (up to 25) which decreases the acceptance ratio for periodic loops in this direction. The values for the critical coupling and the critical exponents for both $z=1$ and $z=2$ are shown in Table \ref{table-nn}, they coincide within the error bars. With the values of Table~\ref{table-nn} the scaling relation $2\beta=\nu(d+z-2+\eta)$ leads to $0.8\pm0.2=z+\eta$, with correlation function exponent $\eta$ (small and usually positive). This rules out $z=2$ and one is led to conclude that indeed $z\sim 1$ for the crystalline transition, independent of $n_{0}$. The small value of $\beta$ rules out mean-field behavior for the CDW transition. Insight into the nature of this transition is gained by the following consideration. In the neighborhood of this transition, i.e., far in the superfluid phase, the x,y-components of the currents ${\cal J}$ fluctuate strongly and may be integrated out as Gaussian fluctuations. In other words: one expects the staggered magnetization order parameter to couple to the gapless 4$^{th}$ sound mode of the superfluid. This yields strong long-range interactions in the time direction for the $\tau$-components of the currents ${\cal J}$. It is likely that these long-range interactions are a relevant perturbation and suppress the exponent $\beta$. Finally, we discuss the data for $n_{0}$ = 0, 0.1, and 0.2. In these cases there is only one phase transition, as the Mott-insulating lobes (phase ``MI'' in Fig.~\ref{cbssfig}) do not have any non-trivial crystalline order. Our data are consistent with a transition in the 3D XY universality class for $n_{0}$=0 and a mean-field phase transition for $n_{0}$ = 0.1 and 0.2. \section{Discussion} \label{sec-d} We studied the $T=0$ phase diagram of two-dimensional interacting bosons by various techniques. The combination of the various aspects of these treatments allow us to draw the following conclusions. For on-site interaction we have seen the lobe structure of the phase boundary in the $t-\mu$ or $J-n_{0}$ plane. This structure is exactly periodic in the chemical potential $n_{0}$ in the Quantum-Phase model of Eq.~(\ref{qp}). In mean field or variational treatments of the different models, the lobes are parabolic near their tips. The Monte Carlo results, however, show that the lobes are sharper than predicted by mean-field approximations. This deviation near the tips of the lobes is expected by scaling arguments: due to the additional particle-hole symmetry at the tips of the lobes, the effective dimension is reduced to $d_{eff}=d+z=3$ and enhanced fluctuations destroy part of the ordered (superfluid) phase. This is consistent with the value of the critical exponent $\nu$ obtained from the simulations ($\nu\approx 2/3$) that agrees well with a 3D XY critical point. Similar sharpening of the lobes is seen in a higher-order $t/U$ expansion for the BH model~\cite{fm-bh,fm-ex}. A central point of this paper is the consideration of finite-range interactions. They lead to a richer phase diagram, due to commensurability effects. In particular, charge-density wave or solid order appears in parts of the phase diagram. In our definition the solid phase is characterized by the breaking of the translation symmetry of the underlying lattice, i.e., if a density wave with a wave vector smaller than the lattice is present. Upon inclusion of nearest-neighbor interactions new half-filled lobes around half-integer $n_{0}$ appear in the phase diagram, in which the checkerboard ($\pi,\pi$) density wave is stable. For next-nearest neighbor interactions additional density waves exist in which 1 or 3 out of 4 sites are occupied, around $n_{0}$=1/4 or 3/4. These pure density waves are incompressible Mott-insulators. Surprisingly, each Mott-insulating charge-density wave seems to have a corresponding compressible supersolid phase in which the particular CDW order coexists with off diagonal LRO. For soft-core bosons the supersolid phases appear already for nearest-neighbor interaction. Hard-core bosons, on the other hand, supersolidify only if at least next-nearest neighbor interaction is present. Another difference is the presence of a supersolid at the tip of the half lobe: hard-core bosons only form a supersolid away from half-integer filling, whereas soft-core bosons do so also at half-integer filling. The difference between Bose-Hubbard and Quantum-Phase models is another issue. Our results indicate that, on the mean-field level, the two models are, apart from a trivial rescaling of the hopping matrix element, almost identical for large $\mu$ or $n_{0}$. For small $\mu$ corresponding to a filling of less than one boson per site the mean-field solution for the two models differ essentially and are not described by a simple rescaling of the parameters. Within our mean-field approximation we find that the supersolid phase is suppressed around the solid lobe with exactly half filling (alternating empty and singly-occupied sites). This may be one of the reasons why in a recent numerical simulation of the BH model \cite{gergy} a supersolid phase was not found for exactly half filling. Our combined mean-field and Monte Carlo analysis has shown that the density-wave order transition is not mean-field like. Both the exponents ($\beta\approx 0.21$ and $\nu\approx 0.55$) and the location of the phase boundary deviate considerably from the mean-field results ($\beta= 1/2$ and $\nu= 1/2$), see also Fig.~\ref{cbssfig}. This may be related to the coupling of the CDW order parameter to the gapless $4^{th}$ sound mode in the superfluid and remains a subject for further study. A related issue is the value of the dynamical critical exponent $z$ for this transition. The value $z$=1 used in the simulations gives good scaling. However, other values of $z$ turn out to give reasonable scaling as well and yield identical values for $\beta$, $\nu$ and $K^{\ast}$. It seems that the CDW transition is rather indifferent to the value of $z$ used to scale the systems studied numerically. From the scaling relation $2\beta/\nu=d+z-2+\eta$ we expect the combination $z$+$\eta\approx$ 0.8$\pm$0.2 which would indicate either a small $\eta$ and renormalized $z\approx$ 0.8 or $z\approx 1$ and $\eta$ negative. Further simulations and renormalization group studies are required for solving this problem. In summary we performed an analysis of interacting bosons in 2D within zero temperature lattice models, using and comparing both mean-field theory and exact Quantum Monte Carlo simulations. We obtained the full phase diagram for on-site interaction as well as for nearest-neighbor interactions, in which case our simulations establish the existence of a supersolid phase in which a CDW and superfluidity coexist. {\bf Acknowledgments:} We thank G.~T. Zimanyi for motivation and discussions, and J.~K. Freericks and H. Monien for discussions and their data as plotted in Fig.~\ref{mottfig}. The work was partially supported by 'Sonderforschungsbereich 195' of the DFG and by the Swiss Nationalfonds (AvO). \begin{table} $$ \begin{array}{||c|c|c|c|c||c|c|c|c|c||} \hline\hline n_{0} & z & K^{\ast} & J^{\ast}/U_0 & \nu & n_{0} & z & K^{\ast} & J^{\ast}/U_0 & \nu \\ \hline\hline 0.00 & 1 & 0.886\pm0.003 & 0.152 & 0.69\pm0.06 & 0.25 & 2 & 0.64\pm0.01 & 0.098 & 0.5\pm0.1 \\ \hline 0.05 & 2 & 0.85\pm0.01 & 0.144 & 0.5\pm0.1 & 0.30 & 2 & 0.56\pm0.01 & 0.081 & `` \\ \hline 0.10 & 2 & 0.82\pm0.02 & 0.137 & `` & 0.35 & 2 & 0.47\pm0.01 & 0.062 & `` \\ \hline 0.15 & 2 & 0.77\pm0.02 & 0.126 & `` & 0.40 & 2 & 0.37\pm0.01 & 0.042 & `` \\ \hline 0.20 & 2 & 0.71\pm0.01 & 0.113 & `` & 0.45 & 2 & 0.26\pm0.01 & 0.023 & `` \\ \hline \end{array} $$ \caption{Critical couplings and exponents for on-site interactions} \label{table-ons} \end{table} \begin{table} $$ \begin{array}{||c||c|c|c|c||c|c|c|c|c||} \hline\hline \multicolumn{10}{||c||}{\mbox{the transition for}} \\ \hline \multicolumn{1}{||c||}{}&\multicolumn{4}{|c||}{\mbox{off-diagonal LRO}}& \multicolumn{5}{|c||}{\mbox{diagonal LRO}} \\ \hline n_{0} & z & K^{\ast} & J^{\ast}/U_0 & \nu & z & K^{\ast} & J^{\ast}/U_0 & \nu & \beta \\ \hline\hline 0.5 & 1 & 0.775\pm0.005 & 0.127 & 0.65\pm0.08 & 1 & \; 0.837\pm0.005& 0.141 & \;0.55\pm0.05 & 0.21\pm 0.04 \\ \hline 0.4 & 2 & 0.645\pm0.008 & 0.099 & 0.44\pm0.08 & 2 & 0.749\pm0.006 & 0.122 & 0.5\pm0.1 & \; 0.25 \pm 0.10 \\ \hline 0.4 & & & & & 1 & 0.747\pm0.007 & 0.121 & 0.55\pm0.06 & \; 0.21 \pm 0.08 \\ \hline \hline 0.2 & 2 & 0.446\pm0.005 & 0.057 & 0.5\pm0.1 & \multicolumn{5}{|c||}{\mbox{for comparison:}} \\ \hline 0.1 & 2 & 0.707\pm0.007 & 0.112 &\; 0.49\pm0.11 & \multicolumn{3}{|c|}{\;\;\;\; \mbox{mean-field:}} & 1/2 & 1/2 \\ \hline 0.0 & 1 & \; 0.843\pm0.005& 0.142 & 0.61\pm0.08 & \multicolumn{3}{|c|}{\;\;\;\;\;\;\;\;\;\; \mbox{3D XY: }} & 2/3 & 1/3 \\ \hline\hline \end{array} $$ \caption{Critical couplings and exponents for the different transitions with nearest-neighbor interactions $U_1/U_0=0.2$} \label{table-nn} \end{table}
\section{Introduction} \indent The minimal supersymmetric Standard Model (MSSM) with the particle content of the Standard Model (SM), including their supersymmetric (SUSY) partners, conserves the discrete quantum number $R$, known as $R$ parity. Under this symmetry, the SM particles are even whereas their superpartners are odd. The quantum number of $R$ parity may conveniently be expressed as~\cite{Rsusy,SW,HS,EGJRV}. \begin{equation} \label{m1} R=(-1)^{3B+L+2S} \, , \end{equation} where $B$, $L$, and $S$ are the baryon number, the lepton number, and the spin of a particle, respectively. Evidently, nonconservation of $R$ parity results, in general, in baryon ($\not\!\! B$) and lepton ($\not\!\! L$) number violating terms. However, the usually considered models retain only the lepton-number-violating terms in order to preserve the stability of the proton. In the presence of both, $\not\!\! L$ and $\not\!\! B$ terms, the proton decay might render the model phenomenologically not viable. The explicit breaking of $R$ parity in a SUSY model means also that we introduce additional coupling constants into the Lagrangian, but not additional fields, {\em i.e.}\ the field content of the model is the same as in the MSSM. This property of the model has some interesting consequences. One important implication is that the explicit lepton-number violation allows, in principle, for a mixing of the left-handed neutrinos with the gauginos and higgsinos. As a consequence, such a mixing, if present, will generate neutrino masses. This mechanism to render neutrinos massive is quite distinctive from the standard options, such as the introduction of right-handed neutrinos or the extension of the Higgs sector by adding an exotic Higgs triplet. However, to make the afore-mentioned mixing possible, one should ensure that the sneutrinos acquire vacuum expectation values (vev's). By a careful examination of the scalar potential of the $R$-parity broken SUSY model, one can show that the appearance of non-vanishing vev's for the scalar neutrinos is not always a direct consequence of the theory. In fact, if a term of the form $\varepsilon_i \varepsilon_{ab}\hat{L}^a_i \hat{H}^b_1$ is neglected in the superpotential of the model, where $\hat{L}_i$ are the chiral lepton fields and $\hat{H}_1$ is the super-Higgs fields with hypercharge $Y=1$, then insistence of non-zero vev's of the sneutrino fields can lead, in some cases, to a fine-tuning relation among the original parameters of the scalar potential. For instance, this would happen if we assumed the scalar mass terms $m_{ij}L^{\dagger}_iL_j +$ H.c.\ to be diagonal in $i$, $j$. On the other hand, the retention of the $\varepsilon_i$ terms leads unavoidably to non-zero vev's of the sneutrinos ($w_i$). It is also easy to see that in the presence of soft-SUSY breaking parameters, the $\varepsilon_i$ terms cannot be rotated away by a unitary transformation. This is consistent with the observation that such a term will be generated radiatively, even if one chooses $\varepsilon_i=0$ at the Planck mass scale~\cite{HS}. Therefore, one could argue that models which break the $R$ parity explicitly, in which the $\varepsilon_i$ terms are neglected, may be considered to be not general and hence incomplete. This fact is expected to influence the analytic expressions of the mass and mixing matrices, especially when new parameters enter the theory. Indeed, we will show how the $\varepsilon_i$ terms change the phenomenological predictions of the model under consideration. Models with an explicit mixing term between $\hat{L}_i$ and $\hat{H}_1$ have been considered in the past~\cite{HS}. More recently, such models have been re-examined by paying special attention to CP violation in the scalar potential as well as to neutrino masses \cite{JN,JNCP}. In Refs.~\cite{HS,JN,JNCP}, it was assumed that all individual lepton numbers, $L_i$, are broken. It has been known for some time that lepton-number-violating interaction can erase the existing baryon asymmetry in the universe (BAU) if $\not\!\! L$ interactions are in thermal equilibrium with the $B+L$-violating sphalerons. The most stringent constraints on $w_i$ and $\varepsilon_i$ arise from such considerations. However, it has been realized~\cite{DR} that one can evade the erasure of the BAU if we demand that one individual lepton number is conserved. No constraints on lepton-number-violating couplings will then follow. The most conservative constraints on lepton-number-violating couplings can then be inferred from an analysis of the absence of possible new-physics phenomena in various experiments. In $R$-parity broken models, there are in general two sources that can produce lepton-number- and lepton-flavour-violating interactions. Those that are induced by the $\varepsilon_i$ and $w_i$ parameters and give rise, {\em e.g.}, to tree level off-diagonal $Z$-boson decays, and those that are proportional to the so-called $\lambda$ and $\lambda'$ couplings in the superpotential. The literature pertinent to constraints on the $\lambda$ and $\lambda'$ couplings is immanent~\cite{hinch,BGH,GBDC,HKK,roh}. Here, we will study effects of lepton-number and/or lepton-flavour violation on the ordinary known matter, which can be induced by $W$- and/or $Z$-boson interactions in an $R$-parity broken model. Therefore, one may consider our study complementary to investigations of limits on the $\lambda$ and $\lambda'$ couplings mentioned above. In an $\not\!\! R$ SUSY model, direct limits on $w_i$ are mostly obtained from upper bounds on light neutrino masses, {\em e.g.}\ using the laboratory bound $m_{\nu_{\tau}} < 31$~MeV~\cite{PDG94}. However, we find that better limits may be deduced by the nonobservation of non-SM processes, such as the decay of a muon into three electrons. Moreover, cosmological and astrophysical implications of a massive light neutrino for our model will be discussed in Section~5.7. The paper is organized as follows. In Section 2, we will give a detailed discussion of the $R$-parity broken SUSY model where the emphasis will be put on the scalar potential of the model. Further analytic results on this topic are relegated in Appendix A. Section 3 treats the mixing between the neutralinos and neutrinos as well as that between the charginos and charged leptons. In Section 4, we derive the $W$- and $Z$-boson interaction Lagrangians in the seesaw approximation. In Section 5, we analyze a number of low-energy processes that can be induced by the non-SM couplings present in $\not\!\! R$ SUSY models. We then discuss new constraints that may be derived by laboratory experiments together with constraints coming from cosmology and astrophysics. In addition, we investigate the possibility that our $R$-parity broken SUSY model can explain the KARMEN anomaly. Our numerical results are presented in Section 6. We draw our conclusions in Section 7. \setcounter{equation}{0} \section{The $R$-parity violating SUSY model} \indent In this section, we set up our definition and notation and outline in some more detail the scalar potential of the MSSM with explicit $R$-parity breaking terms. Our main concern will be to argue that the retention of a special term in the superpotential, which is usually not taken into account, leads {\it naturally}, {\em i.e.}\ without any fine tuning problems, to non-zero vev's of the sneutrinos. As a consequence, neutrinos acquire masses through mixing with gauginos and higgsinos and the standard $W$/$Z$-boson interactions with fermions get modified, {\em e.g.}\ one gets tree-level off-diagonal $Z$ decays. The novelty in our approach is the afore-mentioned term in the superpotential whose coupling strength $\varepsilon_i$ enters the $W$/$Z$ interaction Lagrangians. \subsection{Superpotential} \indent We write the full superpotential $W$ as consisting of an $R$-parity conserving part ($W_0$) and $R$-parity violating term ($W_{\not \! R}$), {\em i.e.}\ \begin{equation} \label{m2} W\ =\ W_0+W_{\not \! R}\, . \end{equation} Let then $\hat{L}_i$ ($\hat{E}_i^C$) and $\hat{Q}_i$ ($\hat{U}_i^C $,$ \hat{D}_i^C$) denote the lepton and quark doublets superfields (lepton and quarks $SU(2)$ singlets) with generation index $i$, respectively and let $\hat{H}_{1,\; 2}$ be the super-Higgs fields. With the usual $U(1)_Y$ quantum number assignment, $Y(\hat{L}_i)=-1$, $Y(\hat{E}_i^C)=2$, $Y(\hat{Q}_i)=1/3$, $Y(\hat{D}_i^C)=2/3$, $Y(\hat{U}_i^C)=-4/3$, $Y(\hat{H}_1)=-1$, $Y(\hat{H}_2)=1$, the standard form for $W_0$ is \begin{equation} \label{m3} W_0\ =\ \varepsilon_{ab}\left [h_{ij}\hat{L}_i^a \hat{H}_1^b \hat{E}_j^C + h'_{ij}\hat{Q}_i^a \hat{H}_1^b \hat{D}_j^C + h''_{ij} \hat{Q}_i^a \hat{H}_2^b \hat{U}_j^C + \mu \hat{H}_1^a \hat{H}_2^b \right]\, , \end{equation} where $a$, $b$ are $SU(2)$ group indices. The explicit breaking of $R$ parity can be introduced through $W_{\not \! R}$, which in its most general form is given by~\cite{SW,HS} \begin{equation} \label{m4} W_{\not \! R}\ =\ \varepsilon_{ab}\left( \lambda_{ijk}\hat{L}_i^a \hat{L}_j^b \hat{E}_k^C + \lambda'_{ijk}\hat{L}_i^a \hat{Q}_j^b \hat{D}_k^C + \varepsilon_i \hat{L}_i^a \hat{H}_2^b \right) + \lambda''_{ijk} \hat{U}_i^C \hat{D}_j^C \hat{D}_k^C \, . \end{equation} Unlike the MSSM, the $R$-parity broken SUSY model allows for explicitly broken lepton ($\not \! \! L$) and baryon ($\not \! \! B$) number interactions. To be more precise, the terms in Eq.~(\ref{m3}) proportional to $\lambda_{ijk}=-\lambda_{jik}$, $\lambda'_{ijk}$ and $\varepsilon_i$ violate lepton number, whereas the baryon number is explicitly broken by the $\lambda''_{ijk}$-term ($\lambda''_{ijk}=-\lambda''_{ikj}$). The presence of both $\not\!\! B$- and $\not\!\! L$-type of terms in the Lagrangian leads to unsuppressed proton decay. Therefore, at the most, we can retain either the $L$-violating or the $B$-violating terms in (\ref{m3}). To account for this fact, hereafter we will set $\lambda''_{ijk}=0$. Let us now comment on the term $\varepsilon_{ab} \varepsilon_i \hat{L}_i^a \hat{H}_2^b$ which is usually not taken into account in $W_{\not \! R}$ by using the argument of field redefinitions to rotate away such bilinears. It has been discussed in some detail in~\cite{JN,JNCP} that this argument may not be valid once we add to the Lagrangian soft-SUSY breaking terms. Indeed, we are unable to absorb the $\varepsilon_i$ terms by using an orthogonal transformation of the $\hat{H}_1$ and $\hat{L}_i$ fields. The omission of such terms is therefore not justified. We note here that exactly these terms are in principle responsible for a non-zero vev's of the sneutrinos. The $\varepsilon_i$ terms force the vev's of the sneutrino fields to assume non-zero values. We will return to this point while discussing the Higgs potential. The $\lambda$-terms in $W_{\not \! R}$ lead to the interaction Lagrangian \begin{equation} \label{Llambda} {\cal L}_{\lambda}\ =\ \lambda_{ijk}\biggl[\tilde{\nu}_{iL} \bar{e}_{jR}e_{jL}+ \tilde{e}_{jL}\bar{e}_{kR}\nu_{iL}+\tilde{e}^*_{kR} \bar{\nu}_{iL}^Ce_{jL} -(i \leftrightarrow j)\biggr]\ +\ \mbox{H.c.}\, , \end{equation} where $e_i$ denote charged leptons enumerated by generation index $i$ and tilded symbols denote as usual the superpartners. Correspondingly, the $\lambda'$ interaction Lagrangian reads \begin{eqnarray} \label{Llambda'} {\cal L}_{\lambda'}&=&\lambda'_{ijk}\biggl( \tilde{\nu}_{iL} \bar{d}_{kR}d_{iL} +\tilde{d}_{jL} \bar{d}_{kR}\nu_{iL}+\tilde{d}^*_{kR}\bar{\nu}_{iL}^Cd_{jL} - \tilde{e}_{iL}\bar{d}_{kR}u_{jL} - \tilde{u}_{jL}\bar{d}_{kR}e_{iL}\nonumber\\ && -\tilde{d}^*_{kR}\bar{e}_{iL}^Cu_{jL} \biggr)\ +\ \mbox{H.c.}\, , \end{eqnarray} where $d_i$ ($u_i$) are down-type (up-type) quarks. Since the main concern of the present paper will be to constrain the lepton-number-violating couplings, it is worth discussing briefly constraints on the $L$-violating couplings which come from baryogenesis. It has been known for some time that non-perturbative anomalous $B$-violating interactions of the electroweak theory can wash out the baryon asymmetry generated initially at the GUT scale. This gives rise to severe constraints on lepton-number-violating couplings, such as the trilinear couplings $\lambda_{ijk}$ and $\lambda'_{ijk}$. In this context, it has been noticed that to evade such limits is sufficient to have one individual lepton number, $L_i$, conserved~\cite{DR}. The latter will be assumed throughout this work. In particular, we have \begin{eqnarray} \label{m7} &&\lambda_{ijk}=0,\quad \mbox{for} \quad i\neq j \neq k\, , \nonumber \\ &&\lambda_{iki} \neq 0,\quad \mbox{for}\quad \not\! \! L_k \,\,\, L_i\, , \nonumber \\ &&\lambda'_{ijk} \neq 0,\quad \mbox{for }\quad \not\! \! L_i\, , \end{eqnarray} where~$\not\! \! L_i$ ($L_i$) indicates which lepton number is violated (conserved). Choosing $L_i$ as the conserved lepton number, we then have four independent $\lambda$ couplings. For instance, conserving $L_e$, we are left with non-zero $\lambda_{121}$, $\lambda_{131}$, $\lambda_{232}$, and $\lambda_{323}$. This conservation of a separate lepton number, {\em i.e.}\ $\Delta L_e=0$, has profound consequences for exotic lepton-number-violating processes, which can, in principle, occur at the tree level in an $R$-parity broken SUSY model. If one $L_i$ is strictly conserved, then no tree-level $\lambda$-dependent interaction can contribute, {\em e.g.}, to the process $\mu \to eee$. The latter would proceed via a sneutrino mediated diagram only if {\em all} individual lepton numbers were broken. If $\Delta L_i=0$ for only one lepton number $L_i$, the bound on the $\lambda$ couplings derived via $\mu \to eee$ in \cite{hinch} does not further apply. The reason is that the process $\mu \to eee$ is still possible at the tree level, however through a diagram containing off-diagonal $Z$ couplings (see discussion in Section~5.1). In the subsequent section, we will also address the question whether having $\Delta L_i=0$ for one lepton number $L_i$ at the level of Lagrangian, this particular symmetry gets broken spontaneously through a vev of the sneutrino field, $w_i$, with the very same flavour index $i$. Indeed, we will argue that this is not the case, when the couplings $\varepsilon_i$ are taken into account. Laboratory constraints on $\lambda$ and $\lambda'$ couplings have been put in~\cite{BGH} and~\cite{GBDC,HKK}. The detectability of possible direct $R$-parity-violating signals through the $\lambda$ and $\lambda'$ couplings at the CERN Large Electron Positron (LEP) collider, planned to operate at 200-GeV centre of mass energies (LEP-2), has been discussed in~\cite{roh}. We note here that there are low-energy processes to which both type of $\not\!\! L$ interactions ---those induced by the trilinear $\not\!\! R$ $\lambda$- and $\lambda'$-dependent couplings and those emanating from non-zero $\varepsilon_i$ and $w_i$ parameters--- will contribute. We will see that a combined analysis is not necessary, {\em i.e.}\ deriving limits first on $\varepsilon_i$ and $w_i$ from, say, $\mu \to eee$, and then applying the so obtained results to put constraints on the $\lambda$ and $\lambda'$ couplings. \subsection{The scalar potential} \indent With the superpotential given, it is straightforward to construct the scalar potential of an $R$-parity broken SUSY model~\cite{Rsusy}. Using the convention that symbols without a hat, say $A$, are the spin-zero content of a chiral superfield $\hat{A}$, we first write down the relevant soft-SUSY breaking terms \begin{eqnarray} \label{m8} V_{soft}&=&m_1^2 H_1^{\dagger}H_1 + m_2^2 H_2^{\dagger} H_2 +(m_{L_{ij}}^2L_i^{\dagger}L_j\ +\ \mbox{H.c.})\nonumber \\ &&-(m_{12}^2\varepsilon_{ab}H_1^a H_2^b\ +\ \mbox{H.c.}) + (\kappa'_i \varepsilon_{ab} H_2^a L_i^b\ +\ \mbox{H.c.}) \nonumber \\ &&+ (\mu^2_{+ij}E_i^* E_j\ +\ \mbox{H.c.})+[\mu'_{ij}(H_1^{\dagger}L_i)E_j\ +\ \mbox{H.c.}] \nonumber \\ &&+ (\kappa'_{ijk}\varepsilon_{ab}L^a_j L^b_j E_k\ +\ \mbox{H.c.})\, . \end{eqnarray} In Eq.~(\ref{m8}), $E_i$ are the positively charged scalar singlet fields. The full scalar potential (without squarks) can be written as the sum of five terms, \begin{equation} \label{m9} V_{Scalar}\ =\ V^{2H} + V^L + V^{\not L} + V_+^L + V_+^{\not L}\, , \end{equation} where $V^{2H}$ is the usual Higgs potential of the MSSM, $V^L$ and $V ^{\not L}$ contain the slepton doublets (both $L$-conserving and $L$-violating parts, the latter connected with $R$-parity breaking couplings), and finally $V_+^L$ and $V_+^{\not L}$ refers to the part of the potential which contains the charged scalars $E_i$. The standard MSSM potential, $V^{2H}$, reads \begin{eqnarray} \label{m10} V^{2H}&=&\mu_1^2\phi_1^{\dagger}\phi_1 + \mu_2^2 \phi_2^{\dagger} \phi_2 +{1 \over 2}\lambda_1(\phi_1^{\dagger}\phi_1)^2 + {1 \over 2}\lambda_1(\phi_1^{\dagger} \phi_2)^2 \nonumber \\ &&+\lambda_3 (\phi_1^{\dagger}\phi_1)(\phi_2^{\dagger}\phi_2) -(\lambda_3 + \lambda_1)(\phi_1^{\dagger}\phi_2)(\phi_2^{\dagger}\phi_1) \nonumber \\ &&+ \lambda_6 (\phi_1^{\dagger}\phi_2) + \lambda_6^*(\phi_2^{\dagger}\phi_1)\, , \end{eqnarray} where we have used the notation $\varphi_i \equiv L_i$, $\phi_2 \equiv H_2$, and $\phi_1 \equiv -i\tau_2 H_1^*$ [$\tau_2$ being the Pauli matrix, ($i\tau_2)_{ab}=\varepsilon_{ab}$]. The new parameters in Eq.~(\ref{m10}), which depend on the couplings entering the superpotential, are the soft-SUSY breaking parameters given in (\ref{m8}), as well as the $SU(2)_L$ coupling constant $g$ and the corresponding $U(1)_Y$ coupling constant $g'$. These parameters are related as follows: \begin{eqnarray} \label{m11} &&\mu_1^2=m_1^2+\vert \mu \vert^2,\; \; \; \mu_2^2=m_2^2 +\vert \mu \vert^2 + \varepsilon_i \varepsilon_i^*\, , \nonumber \\ &&\lambda_1={1 \over 4}(g^2 + g'^2), \; \; \; \lambda_3={1 \over 2}g^2- \lambda_1,\; \; \; \lambda_6=-m_{12}^2\, . \end{eqnarray} The lepton-number-conserving scalar potential containing the slepton doublet fields $\varphi_i$ reads \begin{eqnarray} \label{m12} V^L &=& (\mu^2_{L_{ij}}\varphi_i^{\dagger}\varphi_j\ +\ \mbox{H.c.}) + {1 \over 2}\lambda_1 (\sum_i \varphi_i^{\dagger}\varphi_i)^2 + \lambda_1(\phi_1^{\dagger}\phi_1)(\varphi_i^{\dagger}\varphi_i) \nonumber \\ &&- \lambda_1(\phi_2^{\dagger}\phi_2)(\varphi_i^{\dagger}\varphi_i) + (\lambda_3 + \lambda_1)(\phi_2^{\dagger}\varphi_i)(\varphi_i^{\dagger} \phi_2) \nonumber \\ &&+ \Big[\kappa_{jk} -(\lambda_3 + \lambda_1)\delta_{jk}\Big](\phi_1^{\dagger} \varphi_k)(\varphi_j^{\dagger}\phi_1) + \kappa_{nmij}(\varphi_i^T \tau_2 \varphi_j)(\varphi_n^T \tau_2 \varphi_m)^{\dagger}\, . \end{eqnarray} Here, we have used the definitions \begin{eqnarray} \label{m13} &&\mu_{L_{ij}}^2=m^2_{L_{ij}}+\varepsilon^*_i \varepsilon_j, \,\,\, \kappa_{jk}=\kappa^*_{kj}\equiv h^*_{ji}h_{ki}, \nonumber \\ &&\kappa_{nmij}=-\kappa_{mnij}=-\kappa_{nmji}=\kappa^*_{ijnm} \equiv \lambda^*_{nmk}\lambda_{ijk}\, . \end{eqnarray} $V^L$ given in Eq.~(\ref{m12}) may be contrasted with the corresponding potential of the MSSM. In the MSSM, the term proportional $\kappa_{nmij}$ is absent and $\mu^2_{L_{ij}}$ should be replaced by the diagonal mass parameters $\mu^2_{L_i}=m^2_{L_i}+\vert \varepsilon_i \vert^2$. Finally, the lepton-number-violating part of the potential is given by \begin{eqnarray} \label{m14} V^{\not L}&=&[i\kappa_i (\phi_1^T \tau_2 \varphi_i)\ +\ \mbox{H.c.}] + [i\kappa'_i (\phi_2^T \tau_2 \varphi_i)\ +\ \mbox{H.c.}] \nonumber \\ &&- i\kappa_{nmj}(\phi_1^{\dagger}\varphi_j)(\varphi_n^T \tau_2 \varphi_m)^{\dagger}\, , \end{eqnarray} with \begin{equation} \label{m15} \kappa_i \equiv \mu^* \varepsilon_i, \,\,\, \kappa_{nmj} \equiv \lambda^*_{nmk}h_{jk}\, , \end{equation} and $\kappa'_j$ is a soft-SUSY breaking parameter from Eq.~(\ref{m8}). The parts $V^{2H}$, $V^L$ and $V^{\not L}$ are sufficient to determine the minimization conditions of the potential. The explicit form of the remaining contributions to $V_{Scalar}$, $V_+^L$ and $V_+^{\not L}$, is given in Appendix A. At the minimum, the fields take the values \begin{equation} \label{m16} \langle \phi_1 \rangle =\left(\begin{array}{c} 0 \\ v_1 \end{array}\right), \,\,\, \langle \phi_2 \rangle =\left(\begin{array}{c} 0 \\ v_2 \end{array}\right), \,\,\, \langle \varphi_i \rangle =\left(\begin{array}{c} w_i \\ 0 \end{array}\right), \end{equation} and the minimization conditions are found to be \begin{eqnarray} \label{m17} \mu_1^2 v^*_1 +\lambda_1 v_1^*(\vert v_1 \vert^2 -\vert v_2 \vert^2 + \sum_k \vert w_k \vert^2)+\lambda_6^* v^*_2 -\sum_k \kappa_k w_k &=& 0\, , \nonumber \\ \mu_2^2 v^*_2 -\lambda_1 v_2^*(\vert v_1 \vert^2 -\vert v_2 \vert^2 + \sum_k \vert w_k \vert^2)+\lambda_6 v^*_1 -\sum_k \kappa'_k w_k &=& 0\, , \nonumber \\ \sum_j \mu^2_{L_{ij}}w^*_j +\lambda_1 w_i^*(\vert v_1 \vert^2 -\vert v_2 \vert^2 +\sum_k \vert w_k \vert^2)- \kappa_i v_1-\kappa'_i v_2 &=& 0\, , \end{eqnarray} where the last equation is valid for every generation index $i$. It is clear from Eq.~(\ref{m17}) that $w_i \neq 0$ unless $\kappa_i=\kappa'_i=0$. To further elucidate this point, let us first consider the case of $\kappa_i$ and $\kappa'_i$ not identical to zero ({\em i.e.}\ the case of the $R$-parity broken SUSY model with $\varepsilon_i \neq 0$), and assume $w_i =0$ for some generation index $i$. {}From Eq.~(\ref{m17}), we then get a `fine-tuning' relation among the original parameters of the potential (no summation convention) \begin{equation} \label{m18} (\mu^2_1 + \mu^2_2)\kappa_i^{*}\kappa'^*_i=\lambda^*_6 \kappa_i^{*2}+ \lambda_6 \kappa'^{*2}_i\, . \end{equation} Analogous to our $\not\!\!R$ model is the limiting case of the MSSM, without explicit $R$-parity breaking, if one sets $\kappa_i=\kappa'_i=0$ and replaces $\mu^2_{L_{ij}}$ by the diagonal coupling $\mu^2_{L_i}$. Again, the attempt to maintain $w_i \neq 0$ yields the `fine-tuning' relation (see also \cite{comel}). \begin{equation} \label{m19} (\mu_1^2 - \mu^2_{L_i})(\mu^2_2 + \mu^2_{L_i})=\vert \lambda_6 \vert^2\, . \end{equation} This demonstrates nicely how, unlike the MSSM, the $R$-parity-breaking couplings $\varepsilon_i$ from Eq.~(\ref{m3}) naturally lead to non-zero vev's for the scalar neutrinos. Had we neglected the bilinear term $\varepsilon_{ab} \varepsilon_i \hat{L}_i^a \hat{H}_2^b$ in Eq.~(\ref{m3}), as done usually, then there would have been no compelling reason to acquire non-zero vev's for the scalar neutrino fields even in the $R$-parity broken case. Moreover, if we conserve an individual lepton number, say $L_i$, in the Lagrangian/superpotential of the $R$-parity broken SUSY model ---among other couplings, this implies that $\varepsilon_i=0$---, this symmetry will not break spontaneously; so, we are free to choose the corresponding $\varphi_i$ to have a vanishing vev. This follows again from Eqs.\ (\ref{m18}) and (\ref{m19}). The spontaneous breaking of CP violation in the case of $V_{Scalar}$, Eq.~(\ref{m9}), has been discussed in~\cite{JNCP}. We supplement these considerations on CP properties by giving below all conditions necessary to restore CP conservation in $V^{2H}+V^L+V^{\not L}$. Denoting $\eta_{{}_1}$, $\eta_{{}_2}$ and $\eta_{{}_{L_i}}$ the CP phases of the fields $\phi_1$, $\phi_2$, and $\varphi_i$, respectively, we find from the requirement of CP conservation that (no summation convention below) \begin{eqnarray} \label{m20} &&\lambda_6^* \eta_{{}_1} \eta_{{}_2}^*=\lambda_6\, , \nonumber \\ &&\kappa_{jk}=\kappa^*_{jk}\eta_{{}_{L_i}}\eta^*_{{}_{L_k}}\, , \nonumber \\ &&\kappa_{nmjk}=\kappa^*_{nmjk}\eta_{{}_{L_n}} \eta_{{}_{L_m}}\eta^*_{{}_{L_i}}\eta^*_{{}_{L_j}}\, , \nonumber \\ &&\kappa^*_i \eta^*_{{}_1} \eta^*_{{}_{L_i}}=\kappa_i\, , \nonumber \\ &&\kappa'^*_j \eta^*_{{}_2} \eta^*_{{}_{L_j}}=\kappa'_j\, , \nonumber \\ &&\kappa^*_{nmj}\eta_{{}_1} \eta^*_{{}_{L_j}} \eta_{{}_{L_n}}\eta_{{}_{L_m}}=\kappa_{nmj}\, . \end{eqnarray} {}From Eq.\ (\ref{m20}), one can derive over twenty conditions for CP conservation which in contrast to (\ref{m20}) do not involve the CP phases $\eta$'s. However, it is obvious that not all such conditions are independent. For instance, a set of independent conditions that follows from (\ref{m20}) is given by \begin{eqnarray} \label{m21} &&\Im m(\lambda_6 \kappa_i \kappa'^*_j \kappa_{ij})=0\, , \nonumber \\ &&\Im m(\kappa_i \kappa^*_j \kappa_{ij})=0\, , \nonumber \\ &&\Im m(\kappa'_i \kappa'^*_j \kappa_{ij})=0\, , \nonumber \\ &&\Im m(\kappa^*_{ni}\kappa^*_{mj}\kappa_{nmij})=0\, , \nonumber \\ &&\Im m(\kappa^*_{n}\kappa_{mj}\kappa^*_{nmj})=0\, , \end{eqnarray} where again no summation convention has been used. After spontaneous symmetry breaking, the first equation in the set (\ref{m21}) for $i=j$, {\em i.e.}\ $\Im m(\lambda_6 \kappa_i \kappa^{'*}_i)=0$, translates into the following three equivalent conditions: \begin{eqnarray} \label{m22} &&\Im m(\kappa_i w_i v_1)=0\, , \nonumber \\ &&\Im m(\kappa'_i w_i v_2)=0\, , \nonumber \\ &&\Im m(\lambda_6 v_1^* v_2)=0\, . \end{eqnarray} For the case at hand, it is, however, possible to have complex vev's such that CP gets broken spontaneously. For further details on this issue, the reader is referred to~\cite{JNCP}. Some remarks on the vev's of sneutrinos, $w_i$, are in order. One should notice that through the kinetic term \begin{displaymath} \sum_i \left(D_{\mu}{\cal S}_i \right)^{\dagger}\left(D^{\mu}{\cal S}_i\right)\, , \end{displaymath} where $D_{\mu}$ is the covariant derivative and ${\cal S}_i$ are the scalar fields in the theory, the sneutrino vev's contribute to the gauge bosons masses. In this way, the SM vev is obtained by \begin{equation} \label{mex2} v \equiv \sqrt{v_1^2 + v_2^2 + \sum_i w_i^2}\ =\ \frac{2M_W}{g}\, . \end{equation} As a consequence, $w_i$ and the angle $\beta$ defined by \begin{equation} \label{mex3} \tan \beta = { v_1 \over v_2}\, , \end{equation} for real $v_i$, may be regarded as free parameters of the theory, while $v_i$ are not free any longer, but determined by \begin{eqnarray} \label{mex4} &&v_1 = \sin \beta \sqrt{v^2 -\sum_i w^2_i}\, , \nonumber \\ &&v_2 = \cos \beta \sqrt{v^2 -\sum_i w^2_i}\, . \end{eqnarray} Evidently, the vev's of the scalar neutrinos, $w_i$, cannot have arbitrarily large values, but they are bounded from above, as can be readily seen from Eqs.~(\ref{mex2}) and (\ref{mex4}). \setcounter{equation}{0} \section{Mass matrices} \indent In this section, we will present the mass matrices of the neutralino/neutrino as well as those of the chargino/charged lepton states. Since the lepton number is explicitly broken, the neutrinos will mix with the neutralinos to give the neutrinos mass. A natural seesaw mechanism emerges, in which $\mu$, $v_1$, $v_2$, and the gaugino mass parameters $M$ and $M'$ act as the heavy scales, and the lepton-number-breaking couplings $\varepsilon_i$ together with the sneutrino vev's $w_i$ constitute the light Dirac components of the seesaw matrix. It will turn out that only one neutrino becomes massive through this mechanism at the tree level. These considerations are relevant for putting limits on the lepton-number-breaking parameters. In particular, one can already infer constraints on those parameters from the $\tau$-neutrino mass. Furthermore, the neutralino--neutrino or chargino--charged lepton mixing will enter the interaction Lagrangians of $W$ and $Z$, giving rise to non-SM processes, through which the new parameters can also be constrained. \subsection{Neutralino--neutrino mixing} \indent In general, there are two mechanisms that can give rise to neutrino masses in the Born approximation. For example, one possibility is to give masses to the left-handed neutrinos through the vev of an exotic Higgs field which transform under $SU(2)_L$ as a triplet. The other mechanism requires, in general, the mixing of the left-handed neutrinos with other neutral fields of the theory. The latter are usually taken to be the right-handed neutrinos, introducing hereby additional fields in the theory. In our minimal $R$-parity broken SUSY model, in which right-handed neutrinos are absent, the r\^ole of the new neutral fields required for the afore-mentioned mixing will be assumed by the gauginos and higgsinos. In two component notation, let $\Psi'$ denote the column vector of neutrinos and neutralinos \begin{equation} \label{m23} \Psi^{'T}_0=(\psi^1_{L_1},\,\, \psi^1_{L_2},\,\, \psi^1_{L_3},\,\, -i\lambda',\,\, -i\lambda_3,\,\, \psi^1_{H_1},\,\, \psi^2_{H_2})\, , \end{equation} where $\psi^1_{L_i}$ are the neutrino fields ---the upper index indicates the component of the doublet---, $-i\lambda'$ and $-i\lambda_3$ are the unmixed photino and gaugino states, respectively, and the last two entries refer to the two higgsino fields. In the Weyl basis, the Lagrangian describing the neutralino/neutrino masses is then given by \begin{equation} \label{m24} {\cal L}^{\chi^0}_{mass}=-{1 \over 2}\Psi_0^{'T}{\cal M}_0 \Psi'_0\ +\ \mbox{H.c.}\, , \end{equation} where the mass matrix has the general seesaw-type structure \begin{equation} \label{m25} {\cal M}_0=\left(\begin{array}{cc} 0 & m \\ m^T & M_4 \end{array}\right) . \end{equation} Here, the sub-matrix $m$ is the following $3 \times 4$ dimensional matrix: \begin{equation} \label{m26} m=\left(\begin{array}{cccc} -{1 \over 2}g' w_e & {1 \over 2}g w_e & 0 & -\varepsilon_e \\ -{1 \over 2}g' w_\mu & {1 \over 2}g w_\mu & 0 & -\varepsilon_\mu \\ -{1 \over 2}g' w_\tau & {1 \over 2}g w_\tau & 0 & -\varepsilon_\tau \end{array}\right) . \end{equation} In Eq.~(\ref{m25}), $M_4$ is the usual $4 \times 4$ dimensional neutralino mass matrix of the MSSM, which has the form \begin{equation} \label{m27} M_4=\left(\begin{array}{cccc} cM & 0 & -{1 \over 2}g'v_1 & {1 \over 2}g'v_2 \\ 0 & M & {1 \over 2}gv_1 & -{1 \over 2}gv_2 \\ -{1 \over 2}g'v_1 & {1 \over 2}gv_1 & 0 & - \mu \\ {1 \over 2}g'v_2 & -{1 \over 2}gv_2 & - \mu & 0 \end{array}\right) , \end{equation} where $M$ is the common gaugino mass parameter and $\displaystyle{c={5 g'^2 \over 3g^2 }\simeq 0.5}$. The seesaw hierarchy is now evident, when constraints on neutralino masses and upper limits on lepton-number-violating couplings will be considered in Section~5. We will then find that $(M_4)_{ij} \gg m_{kl}$ in agreement with experimental constraints on neutralino and neutrino masses. We can utilize this posterior fact to calculate the diagonalization of ${\cal M}_0$ in an approximate way in terms of the small matrix-valued quantity defined as \begin{equation}\label{xi} \xi=mM_4^{-1}\, . \end{equation} Parenthetically, we wish to draw the reader's attention to one exact result in connection with the diagonalization of ${\cal M}_0$. Because of the different hypercharge assignments of the two higgsinos and the absence of light-neutrino masses at the tree level, the first three lines together with the last line in ${\cal M}_0$ are not linearly independent. As an immediate consequence of the latter, two neutrino masses are exactly zero in the Born approximation \cite{JN}. Let us now define the mass eigenstates $\Psi_0$ by the rotation \begin{eqnarray} \label{m28} \Psi_{0i}=\Xi_{ij} \Psi'_{0j}\, , \nonumber \\ \Xi^*{\cal M}_0 \Xi^{\dagger} =\widehat{{\cal M}}_0\, , \end{eqnarray} where $\widehat{{\cal M}}_0$ is the diagonal matrix with neutrino/neutralino masses as elements. To leading order in $\xi$ expansion, the approximate form of $\Xi^*$ is readily estimated to be \begin{equation} \label{m30} \Xi^*=\left(\begin{array}{cc} V_\nu^T & 0 \\ 0 & N^* \end{array}\right) \left(\begin{array}{cc} 1 -{1 \over 2}\xi \xi^{\dagger} & -\xi \\ \xi^{\dagger} & 1 -{1 \over 2}\xi^\dagger \xi \end{array}\right), \end{equation} where the second matrix block-diagonalizes ${\cal M}_0$ to the form $\mbox{diag}(m_{eff}, M_4)$ with \begin{equation} \label{m31} m_{eff}=-m\; M_4^{-1}\; m^T = {cg^2+ g'^2 \over D}\, \left(\begin{array}{ccc} \Lambda_e^2 & \Lambda_e \Lambda_\mu & \Lambda_e \Lambda_\tau \\ \Lambda_e \Lambda_\mu & \Lambda_\mu^2 & \Lambda_\mu \Lambda_\tau \\ \Lambda_e \Lambda_\tau & \Lambda_\mu \Lambda_\tau & \Lambda_\tau^2 \end{array}\right). \end{equation} The quantities $\Lambda_i$ and $D$ newly introduced are defined as follows: \begin{equation} \label{31} \vec{\Lambda}\equiv \mu \vec{w} - v_1 \vec{\varepsilon}\, , \end{equation} and \begin{equation} \label{32} D \equiv 4{det M_4 \over M}=2\mu \left[-2cM\mu + v_1v_2 \left(cg^2 + g'^2 \right)\right] . \end{equation} The sub-matrices $N$ and $V_\nu$ in Eq.~(\ref{m30}) diagonalize $M_4$ and $m_{eff}$ in the following way: \begin{equation} \label{33} N^* M_4 N^{\dagger}=\mbox{diag}(m_{\tilde{\chi}^0_i})\, , \end{equation} where $m_{\tilde{\chi}^0_i}$ are the heavy neutralino masses only. For the diagonalization of $M_4$, we have retained the notation and convention of Ref.~\cite{HK}. For the neutrino case, we obtain \begin{equation} \label{m34} V^T_{\nu}\; m_{eff}\; V_{\nu} = \mbox{diag}(0, \; 0, \; m_{\nu}) \, , \end{equation} where the only non-zero neutrino mass is given by \begin{equation} \label{m35} m_{\nu}=tr(m_{eff})={cg^2+ g'^2 \over D}\, \vert \vec{\Lambda} \vert^2\, . \end{equation} Furthermore, an analytic calculation of the rotation matrix $V_{\nu}$ gives \cite{JN} \begin{equation} \label{m36} V_{\nu}=\left(\begin{array}{ccc} \cos \theta_{13} & 0 & -\sin \theta_{13} \\ \sin \theta_{23}\sin \theta_{13} & \cos \theta_{23} & \sin \theta_{23} \cos \theta_{13} \\ \sin \theta_{13} & \sin \theta_{23} & \cos \theta_{13}\cos \theta_{23} \end{array}\right) , \end{equation} where the mixing angles are expressed through the vector $\vec{\Lambda}$ as follows: \begin{equation} \label{m37} \tan \theta_{13} = -{\Lambda_e \over \sqrt{\Lambda_\mu^2 + \Lambda_\tau^2}}, \;\;\;\;\; \tan \theta_{23} = {\Lambda_\mu \over \Lambda_\tau} \, . \end{equation} In~\cite{JN}, the baryogenesis constraint on all lepton-number-violating couplings were applied, which led to a solution to the solar neutrino puzzle through vacuum oscillations. In that case, the neutrino mass $m_\nu$ came out rather naturally of order $10^{-5}$ eV, while the mixing angle $\theta_{13}$ was predicted to be large, {\em i.e.}\ $\tan\theta_{13} \simeq -1/\sqrt{2}$. Since we can evade the constraints from BAU by conserving one individual lepton number, our scenario regarding the light neutrino mass, $m_\nu$, is quite different. In Section 6, we will discuss some numerical examples of the neutrino mass as well as the resulting constraints on $w_i$, $\varepsilon_i$ together with the constraints emerging from exotic processes. Here we note in passing that appreciable values for $w_i$ and $\varepsilon_i$ in the GeV range result in a tau-neutrino mass of ${\cal O}($MeV) which is still allowed by laboratory constraints. Let us now demonstrate explicitly, by an example, how the $\varepsilon_i$ terms can change some of the phenomenological implications. We choose the following set of parameters: $M=\mu=2M_W$, $\tan\beta =1$, $\varepsilon_{\tau}=w_{\tau}=0$, $w_e=w_{\mu}\equiv w=1$~GeV, $\varepsilon_{e}=\varepsilon_{\mu} \equiv \varepsilon$. If we now put $\varepsilon=0$, then $m_{\nu} \simeq 3.5$~MeV. On the other hand, the same soft-SUSY parameters and vev's, but having now $\varepsilon=4w$ instead, give $m_{\nu} \simeq 38.5$~MeV, which already exceeds the laboratory limit on the tau-neutrino mass. \subsection{Chargino--charged lepton mixing} \indent Similar to the case of neutralino--neutrino mixing, the explicit violation of the lepton number allows also for chargino--charged lepton mixing. In two component notation, the mass term takes the form \begin{equation} \label{m38} {\cal L}^{\chi^+}_{mass}= -\zeta^{'T} {\cal M}_+\, \omega'\ +\ \mbox{H.c.}\, , \end{equation} where in the vector $\zeta'$, we gather the lower components of a charged Dirac spinor in the Weyl representation, {\em i.e.}\ \begin{equation} \label{m39} \zeta^{'T}=(\psi^2_{L_1},\,\, \psi^2_{L_2},\,\, \psi^2_{L_3},\,\, -i\lambda_-,\,\, \psi^2_{H_1})\, , \end{equation} whereas $\omega'$ contains the upper components \begin{equation} \label{m40} \omega^{'T}=(\psi_{R_1},\,\, \psi_{R_2},\,\, \psi_{R_3},\,\, -i\lambda_+,\,\, \psi^1_{H_2})\, . \end{equation} In order to establish contact between the notation of the MSSM in~\cite{HK} or that of our minimal $\not\!\! R$ model and the usual SM notation, we note that the charged leptons are represented by their charged conjugate fields, {\em i.e.} \begin{displaymath} l_i^C\ =\ \left( \begin{array}{c} \psi_{R_i} \\ \bar{\psi}_{L_i}^2 \end{array} \right). \end{displaymath} In this basis, the chargino/charged-lepton mass matrix ${\cal M}_+$ appearing in Eq.~(\ref{m38}) may be written down as \begin{equation} \label{m41} {\cal M}_+=\left(\begin{array}{cc} M_l & E \\ E' & S \end{array}\right) , \end{equation} where $S$ is the usual MSSM chargino mass matrix given by \begin{equation} \label{m42} S=\left(\begin{array}{cc} M & \frac{1}{\sqrt{2}} g v_2 \\ \frac{1}{\sqrt{2}} g v_1 & \mu \end{array}\right) . \end{equation} The sub-matrices $E$ and $E'$ which give rise to chargino--charged lepton mixing are defined as follows: \begin{equation} \label{m43} \displaystyle{E=\left(\begin{array}{cc} {1\over \sqrt{2}} g w_e & \varepsilon_e \\ {1\over \sqrt{2}} g w_\mu & \varepsilon_{\mu} \\ {1\over \sqrt{2}} g w_\tau & \varepsilon_{\tau} \end{array}\right)} , \end{equation} and \begin{equation} \label{m44} E'=\left(\begin{array}{ccc} 0 & 0 & 0 \\ \Upsilon_e & \Upsilon_{\mu} & \Upsilon_{\tau} \end{array}\right) , \end{equation} where $\displaystyle{\Upsilon_l \sim {m_l \over v_1} w_l}$ and $m_l$ are the lepton masses. For our numerical purposes, we will assume that $M_l$ is a diagonal matrix whose elements can be identified, to a high accuracy, with the physical lepton masses $m_i$. In addition, we can neglect the elements of $E'$ as compared to the other entries in Eq.~(\ref{m41}). Therefore, we will be working in the approximation $E'=0$. Let us now express the mass eigenstates $\zeta$ and $\omega$ in terms of the states $\zeta'$ and $\omega'$ via the unitary transformations \begin{equation} \label{m45} \zeta_i=\Sigma_{ij}\zeta'_j, \,\,\, \omega=\Omega_{ij}\omega'_j\, . \end{equation} The bi-diagonalization leads then to the diagonal matrix $\widehat{{\cal M}}_+$ whose elements are the chargino and lepton masses \begin{equation} \label{m46} \Sigma^* {\cal M}_+ \Omega^{\dagger}=\widehat{{\cal M}}_+\, . \end{equation} Proceeding now as in the case of the neutralino--neutrino mixing, we carry out an approximate diagonalization for ${\cal M}_+$. In this way, the expansion parameters are found to be \begin{eqnarray} \label{m47} &&\xi_{{}_{L}}^*=ES^{-1} \, ,\nonumber \\ &&\xi_{{}_{R}}^*=M_l^{\dagger}ES^{-1}(S^{-1})^T =M_l^{\dagger}\xi_{{}_{L}}^*(S^{-1})^T\, . \end{eqnarray} Note that $\xi_{{}_{R}} \sim \xi_{{}_{L}} m_l/M$. To leading order in $\xi_{{}_L}$ and $\xi_{{}_R}$, the rotation matrices are written down as \begin{equation} \label{m48} \Sigma^*=\left(\begin{array}{cc} V_L & 0 \\ 0 & U^* \end{array}\right)\left(\begin{array}{cc} 1 -{1 \over 2} \xi_{{}_{L}}^*\xi_{{}_{L}}^T & -\xi_{{}_{L}}^* \\ \xi_{{}_{L}}^T & 1 -{1 \over 2} \xi_{{}_{L}}^T\xi_{{}_{L}}^* \end{array}\right) , \end{equation} and \begin{equation} \label{m49} \Omega^{\dagger}= \left(\begin{array}{cc} 1 -{1 \over 2} \xi_{{}_{R}}^*\xi_{{}_{R}}^T & \xi_{{}_{R}}^* \\ -\xi_{{}_{R}}^T & 1 -{1 \over 2} \xi_{{}_{R}}^T\xi_{{}_{R}}^* \end{array}\right) \left(\begin{array}{cc} V_R^{\dagger} & 0 \\ 0 & V^{\dagger} \end{array}\right). \end{equation} Adopting the convention of~\cite{HK} for the matrices, which also appear in the MSSM, we have \begin{eqnarray} \label{m50} && U^*SV^{\dagger}=\widehat{S}\, , \nonumber \\ && V_L M_l V_R^{\dagger} =\widehat{M}_l\, , \end{eqnarray} where, as before, the hatted matrices are diagonal. \setcounter{equation}{0} \section{The $W$- and $Z$-boson interaction Lagrangians} \indent In this section, we will derive the interaction Lagrangians of $Z$ and $W$ bosons with neutralinos/neutrinos and charginos/charged leptons. We will first present general expressions and, subsequently, use the analytic results of the approximate diagonalization of the mass matrices, given in the preceeding section, to calculate the mixing matrices in the first order approximation. Here and in the following, because of the mixing, we collectively call $\Psi_0$ all the neutralinos, with neutrinos being the light neutralinos, and $\zeta$, $\omega$ all the charginos, where the charged leptons are the light charginos. \subsection{General expressions} \indent Starting from two component notation and defining for convenience the matrix \begin{equation} \label{m51} T^Z=\mbox{diag}(1,\, 1,\, 1,\, 0,\, 0,\, 1,\, -1)\, , \end{equation} the interaction Lagrangian of $Z$ with neutralinos reads \begin{equation} \label{m52} {\cal L}^{Z\chi^0\chi^0}_{int}\ =\ -{g \over 2\cos\theta_w} Z^{\mu}\bar{\Psi}'_{0i} T^Z_{ij}\bar{\sigma}_{\mu}\Psi'_{0j} \, . \end{equation} After replacing the weak eigenstates $\Psi'_{0i}$ by four component Majorana mass eigenstates $\chi^0_i$ in Eq.~(\ref{m52}), we obtain \begin{equation} \label{Znunu} {\cal L}^{Z\chi^0\chi^0}_{int}\ =\ - \frac{g}{4\cos\theta_w}\, Z^\mu\, \bar{\chi}^0_i \gamma_\mu\Big( i \Im m \tilde{C}_{ij}\ -\ \gamma_5 \Re e \tilde{C}_{ij}\Big) \chi^0_j\, , \end{equation} where \begin{equation} \label{m54} \tilde{C}_{ij}\ =\ (\tilde{C}^{\dagger})_{ij}\ =\ (\Xi T^Z \Xi^{\dagger})_{ij} \, . \end{equation} It is easy to check that Eq.~(\ref{Znunu}) reproduces the $Z$-neutralino-neutralino interaction of the MSSM, when the leptonic $\not\!\! R$ admixture is neglected. To calculate the $Z$-chargino-chargino coupling, we again define two auxiliary matrices \begin{eqnarray} \label{m55} T^Z_R &=& \mbox{diag}(0,\,0,\, 0,\,2,\, 1)\, , \nonumber \\ T^Z_L &=& \mbox{diag}(1,\,1,\, 1,\,2,\, 1)\, , \end{eqnarray} such that ${\cal L}^{Z\bar{\chi}^-\chi^-}_{int}$ takes the form \begin{eqnarray} \label{m56} {\cal L}^{Z\bar{\chi}^-\chi^-}_{int} &=&{g \over 2\cos \theta_w}Z^{\mu}\Big[ \bar{\zeta}'_i(T^Z_L)_{ij}\bar{\sigma}_{\mu}\zeta'_j - \bar{\omega}'_i(T^Z_R)_{ij}\bar{\sigma}_{\mu}\omega'_j \nonumber \\ && +2 \sin^2 \theta_w ( \bar{\omega}'_i \bar{\sigma}_{\mu}\omega'_i -\bar{\zeta}'_i\bar{\sigma}_{\mu}\zeta'_i )\Big] . \end{eqnarray} Denoting by $\chi^{-}_i$ the physical charginos in the four-component Dirac notation pertaining to the definition of Eq.~(\ref{m39}), the above Lagrangian can be written down as \begin{equation}\label{Zll} {\cal L}^{Z\bar{\chi}^-\chi^-}_{int}\ =\ \frac{g}{2\cos\theta_w}\, Z^\mu\, \bar{\chi}^-_i \gamma_\mu \Big( \tilde{A}^L_{ij} \mbox{P}_L\ +\ \tilde{A}^R_{ij} \mbox{P}_R\Big)\chi^-_j\, ,\\ \end{equation} where P$_L$(P$_R)=[1-(+)\gamma_5]/2$ and \begin{eqnarray} \label{m58} \tilde{A}^L_{ij} &=& (\Sigma T^Z_L \Sigma^{\dagger})_{ij} - 2\delta_{ij} \sin^2 \theta_w\, , \nonumber \\ \tilde{A}^R_{ij}&=&(\Omega^* T^Z_R \Omega^T)_{ij} - 2\delta_{ij} \sin^2 \theta_w \, . \end{eqnarray} As done above for the neutral-current interactions, for the charged-current case we first introduce two auxiliary matrices given by \begin{equation} \label{m59} T^L=\left(\begin{array}{ccccccc} 1 & 0 & 0 & 0 & 0 & 0 & 0 \\ 0 & 1 & 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & \sqrt{2} & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 & 1 & 0 \end{array}\right), \, \, \, T^R=\left(\begin{array}{ccccc} 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & -\sqrt{2} & 0 \\ 0 & 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 & 1 \end{array}\right) . \end{equation} In the Weyl weak eigenbasis, we can then write \begin{equation} \label{m60} {\cal L}^{W\chi^-\chi^0}_{int}\ =\ -{ g \over \sqrt{2}} \Big( \bar{\zeta}'_i T^L_{ij}\bar{\sigma}^{\mu}\Psi'_{0j} + \bar{\Psi'}_{0i} T^R_{ij}\bar{\sigma}^{\mu}\omega'_j\Big) W^-_{\mu}\ +\ \mbox{H.c.}\, , \end{equation} and \begin{equation}\label{Wlnu} {\cal L}^{W\chi^-\chi^0}_{int} \ =\ -\frac{g}{\sqrt{2}}\, W^{-\mu}\, \bar{\chi}^-_i \gamma_\mu \Big( \tilde{B}^L_{ij} \mbox{P}_L\ +\ \tilde{B}^R_{ij} \mbox{P}_R\Big) \chi^0_j\ +\ \mbox{H.c.}\, ,\\ \end{equation} in four-component mass eigenbasis notation. The mixing matrices are \begin{eqnarray} \label{m62} \tilde{B}^L_{ij} &=& (\Sigma T^L \Xi^\dagger)_{ij}\, , \nonumber \\ \tilde{B}^R_{ij} &=& -[\Omega^* (T^R)^T \Xi^T ]_{ij}\, . \end{eqnarray} In the next section, we will give analytic approximate expressions for all mixing matrices defined above. \subsection{Mixing matrices} \indent In Section~4.1, we have derived the analytic expressions of the interaction Lagrangians of the $W$ and $Z$ bosons with chargino and neutralino states in the $R$-parity-violating SUSY model. However, the mixings $\tilde{A}^L$, $\tilde{A}^R$, $\tilde{B}^L$, $\tilde{B}^R$, and $\tilde{C}$ that govern these interactions are high dimensional matrices, involving a large number of parameters. Therefore, it is more convenient to find approximative forms for the mixing matrices that will enable us to appreciate the strength of the $Z$- and $W$-boson couplings in the model under consideration. To facilitate our presentation, we first introduce the following auxiliary matrices: \begin{eqnarray} d &=& \mbox{diag}(2,1)\, ,\qquad t_z\ =\ \mbox{diag}(0,0,1,-1)\, ,\nonumber\\ t_{{}_L}& =& \left( \begin{array}{cccc} 0 & \sqrt{2} & 0 & 0 \\ 0 & 0 & 1 & 0 \end{array} \right),\nonumber\\ t_{{}_R}& =& \left( \begin{array}{cccc} 0 & \sqrt{2} & 0 & 0 \\ 0 & 0 & 0 & -1 \end{array} \right). \end{eqnarray} Substituting the unitary mixing matrices of Eqs.\ (\ref{m30}), (\ref{m48}), and (\ref{m49}) into Eqs.~(\ref{m54}), (\ref{m58}), and (\ref{m62}), and neglecting terms of ${\cal O}(\xi_{{}_L}^3,\xi^3 )$ and higher, we obtain \begin{eqnarray} \tilde{A}^L &=& -2\sin^2\theta_w {\mbox{\bf 1}}\ +\ \left( \begin{array}{cc} 1-\tilde{\xi}_{{}_L}(1-d)\tilde{\xi}_{{}_L}^\dagger & \tilde{\xi}_{{}_L}(1-d)U^\dagger\\ U(1-d)\tilde{\xi}^\dagger_{{}_L} & UdU^\dagger \end{array}\right), \label{AL}\\ \tilde{A}^R &=& -2\sin^2\theta_w {\mbox{\bf 1}}\ +\ \left( \begin{array}{cc} 0 & -\tilde{\xi}_{{}_R}dV^T\\ -V^*d\tilde{\xi}_{{}_R}^\dagger & V^*d V^T \end{array}\right), \label{AR}\\ \tilde{B}^L &=& \left( \begin{array}{cc} V^l-\frac{1}{2}\tilde{\xi}_{{}_L}\tilde{\xi}^\dagger_{{}_L}V^l- \frac{1}{2}V^l\tilde{\xi}^*\tilde{\xi}^T + \tilde{\xi}_{{}_L} t_{{}_L} \tilde{\xi}^T & (V^l\tilde{\xi}^* - \tilde{\xi}_{{}_L} t_{{}_L}) N^\dagger \\ U(\tilde{\xi}^\dagger_{{}_L}V^l-t_{{}_L}\tilde{\xi}^T) & Ut_{{}_L}N^\dagger \end{array} \right),\label{BL}\\ \tilde{B}^R &=& \left( \begin{array}{cc} 0 & - \tilde{\xi}_{{}_R} t_{{}_R} N^T \\ -V^* t_{{}_R} \tilde{\xi}^\dagger & V^* t_{{}_R} N^T \end{array} \right),\label{BR}\\ \tilde{C} &=& \left( \begin{array}{cc} 1- \tilde{\xi}^* (1-t_z)\tilde{\xi}^T & \tilde{\xi}^* (1-t_z) N^\dagger \\ N(1-t_z)\tilde{\xi}^T & N t_z N^\dagger \end{array} \right).\label{C} \end{eqnarray} Here, we have defined $\tilde{\xi}_{{}_L} =V_L^*\xi_{{}_L}$, $\tilde{\xi}_{{}_R}=\widehat{M}_l \tilde{\xi}_{{}_L} (S^{-1})^\dagger$, $\tilde{\xi} = V_\nu^T \xi$, and $V^l=V_L^* V_\nu$. Furthermore, the unitary matrices $V_L$, $V_R$, $V_\nu$, $U$, $V$, $N$, together with the mixing matrices $\xi_{{}_L}$, $\xi_{{}_R}$, and $\xi$ are defined in Section~3. In the derivation of Eqs.~(\ref{AR}) and (\ref{BR}), we have also used the fact that $\xi_{{}_R} ={\cal O}( \xi_{{}_L} m_l/M)$. {}From Eqs.~(\ref{AL})--(\ref{C}), it is now easy to see how the $R$-parity-violating couplings to ordinary leptons deviate from the SM vertices. To leading order in $\xi_{{}_L}$ and $\xi$, we find that the interactions of the $W$ and $Z$ bosons with left-handed charged leptons and neutrinos are modified, whereas the corresponding couplings to right-handed charged leptons remain unaffected, having the SM form. \setcounter{equation}{0} \section{Laboratory and cosmological constraints} \indent Our aim is to constrain the parameter space of this $\not \!\! R$ scenario, by taking laboratory and cosmological constraints into account. For this purpose, we will pay special attention to limits derived from low-energy processes and LEP data, such as charged lepton decays of the form $l^-\to l'^-l_1^-l_1^+$, flavour-changing $Z$-boson decays $Z\to l_il_j$, the invisible width of the $Z$ boson, charged-current universality in muon and tau decays, lepton universality at the $Z$ peak, and charged-current universality in pion decays. In this vein, we will report some phenomenological implications of our minimal model that may be relevant to explain the intriguing anomaly found by the KARMEN collaboration~\cite{KARMEN}. In the last section, we will discuss the viability of our model when cosmological constraints are considered, such as the requirement of not washing out the primordial BAU and the absence of large disruptive reheating effects caused by an unstable $\tau$ neutrino with $m_{\nu_\tau}={\cal O}(10)$~MeV. \subsection{{\boldmath $l^-\to l'^-l_1^-l^+_1$}} \indent As has been found in Lagrangian~(\ref{Zll}), the model predicts flavour-changing neutral current (FCNC) $Zll'$ couplings at the tree level. These new $\not\!\! R$ interactions induce $\tau$ and $\mu$ decays into three lighter charged leptons. In this way, we obtain \begin{equation} B(l^-\to l'^-l_1^-l^+_1)\ =\ \frac{\alpha_w^2 m^4_l}{1536\pi M_W^4}\, \frac{m_l}{\Gamma_l} \left( \left|\tilde{A}^L_{ll'}\right|^2 +\left|\tilde{A}^R_{ll'}\right|^2\right)\left(\left|\tilde{A}^L_{l_1l_1} \right|^2 +\left|\tilde{A}^R_{l_1l_1}\right|^2\right), \label{BRlll} \end{equation} where $\alpha_w=g^2/4\pi$ and $\Gamma_l$ is the total width of the decaying charged lepton $l$. The experimental upper limit on the branching ratio of $\mu^- \to e^-e^-e^+$ is given by~\cite{PDG94} \begin{equation} B(\mu^- \to e^-e^-e^+) \ \leq\ 1.0\cdot 10^{-12}\, ,\label{Bexpmueee} \end{equation} at $90\%$ confidence level (CL). Recently, CLEO collaboration~\cite{CLEO} has considerably lowered experimental upper bounds on branching ratios of neutrinoless $\tau$-lepton decays. They have found \begin{eqnarray} B(\tau^-\to e^- e^+ e^- ) &\leq& 3.3\cdot 10^{-6}\, ,\nonumber\\ B(\tau^-\to \mu^- e^+ e^- ) &\leq& 3.4\cdot 10^{-6}\, ,\nonumber\\ B(\tau^-\to e^- \mu^+ \mu^- ) &\leq& 3.6\cdot 10^{-6}\, ,\nonumber\\ B(\tau^-\to \mu^- \mu^+ \mu^- )&\leq& 4.3\cdot 10^{-6}\, ,\label{Bexptaulll} \end{eqnarray} at $90\%$ CL. Theoretical predictions obtained for the observables given in Eqs.~(\ref{Bexpmueee}) and~(\ref{Bexptaulll}) will be discussed in Section 6. \subsection{{\boldmath $Z\to l^-l'^+$} and {\boldmath $Z\to \nu\nu$}} \indent The presence of FCNC $Zll'$ couplings at the tree level will also give rise to flavour-violating $Z$-boson decays at LEP. The theoretical prediction of their branching ratios is determined by \begin{equation} B(Z\to l^-l'^+\ \mbox{or}\ l^+ l'^-)\ =\ \frac{\alpha_w}{12\cos^2\theta_w}\, \frac{M_Z}{\Gamma_Z}\left( \left|\tilde{A}^L_{ll'}\right|^2 +\left|\tilde{A}^R_{ll'}\right|^2 \right),\label{BZlilj} \end{equation} where $\Gamma_Z=2.49$~GeV is the total width of the $Z$ boson measured experimentally~\cite{PDG94}. Furthermore, an analysis of this kind of decays at LEP yields \begin{eqnarray} B( Z\to e^-\mu^+\ \mbox{or}\ e^+\mu^-) &\leq & 6.0\cdot 10^{-6}\, ,\nonumber\\ B( Z\to e^-\tau^+\ \mbox{or}\ e^+\tau^-) &\leq & 1.3\cdot 10^{-5}\, ,\nonumber\\ B( Z\to \tau^-\mu^+\ \mbox{or}\ \tau^+\mu^-) &\leq & 1.9\cdot 10^{-5}\, , \label{BexpZlilj} \end{eqnarray} at $95\%$ CL. In addition, the Lagrangian (\ref{Znunu}) modifies the invisible width of the $Z$ boson through the non-universal and flavour-dependent $Z\nu_i\nu_j$ tree-level couplings. It is then easy to obtain the branching ratio for the total invisible $Z$-boson width, which is assumed to be caused mainly by $Z\to \nu_i\nu_j$ \begin{equation} B(Z\to \nu\bar{\nu})\ =\ \frac{\alpha_w}{24\cos^2\theta_w}\, \frac{M_Z}{\Gamma_Z}\sum\limits_{\nu_i,\nu_j}\, \left|\tilde{C}_{\nu_i \nu_j}\right|^2. \label{BZnunu} \end{equation} On the other hand, an experimental analysis on the $Z$ pole gives~\cite{PDG94} \begin{equation} 1 - \frac{B(Z\to \nu\bar{\nu})}{B_{SM}(Z\to \mbox{invisible} )}\ \leq \ 1.31\cdot 10^{-2}\, ,\label{BexpZnunu} \end{equation} where $B_{SM} (Z\to \mbox{invisible} )$ is the SM prediction for the invisible width of the $Z$ boson. In Section 6, we will analyze the phenomenological impact of the new-physics decay channels mentioned above on restricting our model. \subsection{ Universality violation at the {\em Z} peak} \indent Interesting limits on $R$-parity breaking, nonuniversal, diagonal $Zll$ couplings can be extracted from measurements of lepton universality on the $Z$-boson pole. In order to impose constraints, we will adopt the LEP observable based on leptonic $Z$-boson partial width differences studied in~\cite{BKPS} \begin{equation} U_{br}^{(ll' )}\ =\ \frac{\Gamma(Z\to l^+l^-) - \Gamma (Z\to l'^+l'^-)}{\Gamma(Z\to l^+l^-) + \Gamma (Z\to l'^+l'^-)} \ =\ \frac{ |\tilde{A}^L_{ll}|^2\ -\ |\tilde{A}^L_{l'l'}|^2}{ |\tilde{A}^L_{ll}|^2\ +\ |\tilde{A}^L_{l'l'}|^2}\, ,\label{Ubr} \end{equation} where $l\ne l'$. A combined experimental analysis for the observable $U_{br}$ gives~\cite{PDG94} \begin{eqnarray} |U_{br}^{(ll')}| \ \leq \ 5.0\cdot 10^{-3}\, ,\label{Ubrexp} \end{eqnarray} at 1$\sigma$ level, almost independent of the charged leptons $l$ and $l'$. Another relevant observable involving leptonic asymmetries, which has been analyzed in~\cite{BP}, is \begin{equation} \Delta{\cal A}_{ll'}\ =\ \frac{{\cal A}_l\ -\ {\cal A}_{l'}}{ {\cal A}_l\ +\ {\cal A}_{l'}}\ =\ \left(\frac{1}{{\cal A}_l^{(SM )}}\ -\ 1\right)\, U_{br}^{(ll')}\, ,\label{DAll'} \end{equation} where ${\cal A}_l^{(SM)}=0.14$ is the leptonic asymmetry predicted theoretically in the SM. In the last step of Eqs.~(\ref{Ubr}) and~(\ref{DAll'}), we have used the fact that, to a good approximation, the tree-level coupling of the $Z$ boson to right-handed charged leptons is universal in our minimal $R$-parity violating SUSY model, {\em i.e.}\ $\tilde{A}^R_{ll}=\tilde{A}^R_{l'l'}$ as can be seen from Eq.~(\ref{AR}). Considering the experimental upper bound on $U_{br}$ given in Eq.~(\ref{Ubrexp}), Eq.~(\ref{DAll'}) furnishes the upper limit \begin{equation} \Delta{\cal A}_{ll'} \ \le\ 3.0\cdot 10^{-2}\, , \label{DAUbr} \end{equation} which is slightly below the present experimental sensitivity at LEP~\cite{LEP} [$\Delta{\cal A}^{LEP}_\tau/{\cal A}^{(SM)}_l=0.07$, at 1$\sigma$] and Stanford Linear Collider (SLC)~\cite{SLD} [$\Delta{\cal A}^{SLC}_e/{\cal A}_l^{(SM)}=0.04$, at 1$\sigma$]. It is also interesting to notice that the apparent difference of $\Delta{\cal A}_{\tau e}\simeq - 10\%$ between the measured leptonic asymmetries ${\cal A}^{SLC}_e$ and ${\cal A}^{LEP}_\tau$ cannot be predicted in our $\not\!\!\! R$ model, without invalidating the inequality~(\ref{DAUbr}) at the same time. \subsection{Decays {\boldmath $\mu\to e\nu\nu$} and {\boldmath $\tau\to e\nu\nu$}} \indent Useful constraints can be obtained from possible deviations of charged-current universality in $\tau$-lepton decays. In fact, measures of such deviations can be defined and straightforwardly be calculated as follows: \begin{eqnarray} R_{\tau e} &=& \frac{\Gamma (\tau\to e \nu\bar{\nu})}{ \Gamma (\mu\to e \nu\bar{\nu})}\ =\ R_{\tau e}^{SM}\, \frac{\sum\limits_{\nu_i} \Big[ |\tilde{B}^L_{\tau\nu_i}|^2 + |\tilde{B}^R_{\tau \nu_i}|^2 \Big]}{\sum\limits_{\nu_j} \Big[ |\tilde{B}^L_{\mu\nu_j}|^2 + |\tilde{B}^R_{\mu \nu_j}|^2 \Big]}\, , \label{Rtaue}\\ R_{\tau \mu} &=& \frac{\Gamma (\tau\to \mu \nu\bar{\nu})}{ \Gamma (\mu\to e \nu\bar{\nu})}\ =\ R_{\tau\mu}^{SM}\, \frac{\sum\limits_{\nu_i} \Big[ |\tilde{B}^L_{\tau\nu_i}|^2 + |\tilde{B}^R_{\tau \nu_i}|^2 \Big]}{\sum\limits_{\nu_j} \Big[ |\tilde{B}^L_{e\nu_j}|^2 + |\tilde{B}^R_{e \nu_j}|^2 \Big]}\, .\label{Rtaumu} \end{eqnarray} In Eqs.~(\ref{Rtaue}) and~(\ref{Rtaumu}), the SM contributions to the observables, $R_{\tau e}^{SM}$ and $R_{\tau\mu}^{SM}$, have been factored out. Of course, deviations from the SM values can also be induced by the $\lambda$-dependent interactions in Eq.~(\ref{Llambda}). These observables are used to constrain the couplings $\lambda_{ijk}$ as a function of the mass of the scalar right-handed leptons~\cite{BGH}. To avoid excessive complication, we assume that all $\lambda_{ijk}=0$ and focus our study mainly on the phenomenological consequences originating from the $\varepsilon_i$ terms in the superpotential. Furthermore, experimental limits related to the ratios $R_{\tau e}$ and $R_{\tau\mu}$ may be presented in the following way~\cite{DAB}: \begin{eqnarray} 1\ -\ \frac{R_{\tau e}}{R^{SM}_{\tau e}} &=& 0.040 \pm 0.024\, ,\label{Rexptaue}\\ 1\ -\ \frac{R_{\tau\mu}}{R^{SM}_{\tau\mu}} &=& 0.032 \pm 0.024\, ,\label{Rexptaumu} \end{eqnarray} at 1$\sigma$ level. Constraints obtained from Eqs.~(\ref{Rexptaue}) and~(\ref{Rexptaumu}) on the parameters of our $\not\!\! R$ model will be discussed in Section 6. \subsection{Charged-current universality in pion decays} \indent Complementary to the physical quantities $R_{\tau e}$ and $R_{\tau\mu}$ are the constraints derived from the ratio $R_\pi=\Gamma (\pi\to e\nu )/\Gamma (\pi\to \mu\nu )$ in the $\pi^-$ decays. $R_\pi$ is an observable that measures possible deviations from charged-current universality in the $e-\mu$ system. It is not difficult to obtain \begin{equation} R_\pi\ \ =\ R_\pi^{SM}\, \frac{\sum\limits_{\nu_i} \Big[ |\tilde{B}^L_{e\nu_i}|^2 + |\tilde{B}^R_{e \nu_i}|^2 \Big]}{\sum\limits_{\nu_j} \Big[ |\tilde{B}^L_{\mu\nu_j}|^2 + |\tilde{B}^R_{\mu \nu_j}|^2 \Big]}\, .\label{Rpi} \end{equation} In addition, the 1$\sigma$ experimental bound related to $R_\pi$ is given by~\cite{DAB} \begin{equation} \frac{R_\pi}{R_\pi^{SM}}\ -\ 1 \ =\ 0.003\pm 0.003\, .\label{Rpiexp} \end{equation} It is again worth mentioning that similar deviations of $e-\mu$ universality can arise from the presence of $\lambda'$-dependent couplings through the interaction Lagrangian~(\ref{Llambda'}). In our analysis, we will assume that all $\lambda'_{ijk}=0$. This may also be reflected by the fact that the current experimental lower bound on the half-lifetime of the $^{76}$Ge $0\nu\beta\beta$ decay leads to the tight constraint~\cite{HKK} \begin{equation} \lambda'_{111}\ \le\ 3.9\cdot 10^{-4}\, \left( \frac{m_{\tilde{q}}}{100\ \mbox{GeV}}\right)^2 \left( \frac{m_{\tilde{g}}}{100\ \mbox{GeV}} \right)^{1/2}, \end{equation} where $\tilde{q}$ ($\tilde{g}$) is the scalar quark (gluino). \subsection{KARMEN anomaly} \indent Recently, the KARMEN collaboration, which operates at RAL, has reported an anomaly~\cite{KARMEN} in the time-dependence of decay spectra coming from stopped pions. To account for the KARMEN anomaly, one can make the plausible assumption that a new massive weakly-interacting particle, say $x$, is produced in the pion decays, {\em i.e.}\ $\pi^+\to \mu^+ x$~\cite{KARMEN,BPS}. The mass of this hypothetical particle should be $m_x\simeq 33.9$ MeV, since it should explain the apparent $\sim 2\sigma$ bump present in the time distribution of decaying muon events, which should normally fall off exponentially. This experimental peak occurs with a time delay of 3.6~$\mu$sec after all pulsed pions have promptly decayed. A recent study~\cite{BPS} suggests that the $x$ particle should have similar features with those of a neutrino, but it cannot be the $\nu_\tau$, because $m_{\nu_\tau}< 31$ MeV at 95$\%$ CL~\cite{PDG94}, or another predominantly-isodoublet neutrino, without affecting limits coming from the supernova 1987A. The authors in~\cite{BPS} further advocate that a mainly-sterile neutrino scenario could, in principle, be compatible with all constraints ---both terrestrial and astrophysical---, since the production of $x$ particles both in supernova and in the early universe could then be suppressed. Although in our $\not\!\! R$ model the coupling mixing matrices describing the charged- and neutral-current interactions differ crucially from usual singlet-neutrino scenarios~\cite{ZPC}, the above discussion is still valid and translates into the requirement that one neutralino state, {\em e.g.}\ $\chi$, should be light, having a mass $m_\chi=m_x$. Assuming that the KARMEN anomaly gets resolved by the decay $\chi\to e^-e^+\nu$, we have for the Majorana fermion $\chi$~\cite{BPS} \begin{eqnarray} |\tilde{B}^L_{e\chi}|\, |\tilde{B}^L_{\mu\chi}| &\simeq & 0.6\cdot 10^{-6}\, , \nonumber\\ |\tilde{B}^L_{e\chi}| &\stackrel{\displaystyle <}{\sim} & 2.5\cdot 10^{-4}\, , \nonumber\\ |\tilde{B}^L_{\mu\chi}| &\stackrel{\displaystyle <}{\sim} & 4.5\cdot 10^{-2}\, {}. \label{KARMENmix} \end{eqnarray} The bounds presented in Eq.~(\ref{KARMENmix}) are obtained from a number of phenomenological requirements, such as the absence of a correction to the Michael $\rho$ parameter in $\mu\to e\nu\nu$, negligible decay events in neutrino beams, no anomalous contributions to $\pi\to e\chi$, limits from neutrinoless double-$\beta$ decays, {\em etc}. Because of the large number of parameters existing in our model, it appears not difficult to accommodate the upper limits and relations given in Eq.~(\ref{KARMENmix}). However, the soft-SUSY breaking parameters in our model have to satisfy the following hierarchy scheme: \begin{eqnarray} M\, (=2M') &\stackrel{\displaystyle >}{\sim}& 500\ \mbox{GeV}\, ,\nonumber\\ \mu &\stackrel{\displaystyle <}{\sim}& 30\ \mbox{MeV}\, ,\nonumber\\ \vec{\varepsilon} & \sim & \frac{\mu}{v_1}\vec{w}\, ,\label{KARMENsusy} \end{eqnarray} which is mainly prescribed by the fact that $m_\chi=33.9$~MeV. {}From Eq.~(\ref{KARMENsusy}), we find that only SUSY models with a $\mu$ at the scale of 10 MeV have a chance to account for the KARMEN anomaly. Similar $R$-parity broken SUSY models were also discussed in Ref.~\cite{EGJRV}. Adapting the results of~\cite{EGJRV}, one can estimate that for $\tan\beta=1$ and $w_\tau<60$~GeV, $B(Z\to \chi \chi)\simeq w_\tau^4/(3v^4)<1.\ 10^{-3}$ in compliance with the LEP bound on invisible $Z$-boson decays in Eq.~(\ref{BexpZnunu}). However, such light-$\mu$ scenarios may encounter the known $\mu$ hierarchy problem, where $\mu\sim M_{Pl}$ as derived naively from supergravity. Even though one could invoke the Guidice--Masiero mechanism~\cite{GM} to obtain a value of $\mu$ at the electroweak scale, the small value of $\mu={\cal O}(10)$ MeV would, however, require an additional unnatural suppression of the gravitational couplings in the K\"ahler potential. \subsection{Cosmological and astrophysical constraints} \indent The minimal SUSY model with explicit $R$ nonconservation contains lepton-number violating interactions that can wash out any primordial BAU generated at the GUT scale via the $B+L$-violating sphaleron interactions~\cite{GTH,NM,KRS}, which are in thermal equilibrium above the critical temperature of the electroweak phase transition~\cite{ADD}. Sphalerons generally conserve the individual quantum numbers $B/3-L_i$~\cite{CDEO,DR}. In particular, it has been shown in~\cite{DR} that if only one separate lepton number is preserved in thermal equilibrium ({\em e.g.}\ $L_i$) and finite masses for the charged leptons are taken into account in the analysis of chemical potentials, this is then sufficient to protect any primordial excess in $L_i$, which can be converted later on, via sphalerons, into the observed BAU. For our purposes, we will assume that only one separate lepton number is conserved each time in the full Lagrangian, when low-energy experiments are considered. For definiteness, in our numerical analysis we will consider that either $w_e=\varepsilon_e=0$ or $w_\tau=\varepsilon_\tau=0$. Of course, one can use a complementary restriction and put $w_\mu=\varepsilon_\mu=0$, which, however, will not alter our phenomenological constraints discussed in Sections~5.1--5.5 in an essential way. There is a great number of bounds coming from astrophysics, such as those obtained from the dynamics of red giants and white dwarfs, or the absence of a distorted spectrum of the $2.73^\circ$ K blackbody radiation background~\cite{GGR}. However, we find more worrying the severe limits derived from possible reheating effects of a decaying massive neutral relic with $m_\nu\simeq 10-40$ MeV and especially those obtained from the primordial nucleosynthesis~\cite{EGLNS,EMR}. In particular, $\tau$-neutrino decays with a lifetime bigger than about 1~sec or so may increase the elemental ${}^4$He abundance by making it incompatible with astrophysical observations. Imposing the latter constrain, we find \begin{equation} |\tilde{B}^L_{e\nu_\tau}|^2\ \stackrel{\displaystyle >}{\sim}\ 10^{-4}\, \left(\frac{30\, \mbox{MeV} }{m_{\nu_\tau}}\right)^5\, , \label{Bastr} \end{equation} which is only applicable for $m_{\nu_\tau} \stackrel{\displaystyle >}{\sim} 10-50$~MeV~\cite{EGLNS}. In fact, the bound of Eq.~(\ref{Bastr}) is not so restrictive, since it simply constrains only the mixing-matrix element $V^l_{e\nu_\tau}>10^{-2}$ in Eq.~(\ref{BL}), which is not excluded from solar neutrino oscillation scenarios. In our analysis of laboratory observables, we sum up over all invisible light neutrinos, so the unitary matrix $V^l$ becomes practically redundant. Moreover, the nonobservation of a $\gamma$ ray burst from the Solar Maximum satellite after the supernova 1987A neutrinos were detected may point towards the fact that the $\tau$ neutrino mainly decays inside the supernova core. This leads again to $\nu_\tau$ lifetimes compatible with the approximate inequality of Eq.~(\ref{Bastr}). Even though the predicted supernova luminosity will increase in such a case, an allowed window of scenarios that maximally violate $L_\mu$ and $L_\tau$ may be present in the $\sim 3$ MeV neutrino-sphere~\cite{BBHH}. As has also been pointed out by the authors in Ref.~\cite{BBHH}, there may exist viable cosmological models in which $\nu_\tau$ is stable with a mass of ${\cal O}(10)$~MeV. Such a solution requires an alteration of the standard cosmological picture by, {\em e.g.}, reheating the universe even after inflation to only a few MeV and invoking low-temperature baryogenesis as well~\cite{DH}. Then, the resulting $\nu_\tau$ may not overclose the universe but it can even constitute the cold dark matter. For neutrinos with $m_\nu\stackrel{\displaystyle <}{\sim} 0.1$~MeV, the cosmological bound regarding their lifetimes, $\tau_\nu$, is different. In fact, $\tau_\nu$ should not be larger the age of the universe, {\em i.e.}\ $\tau_\nu \stackrel{\displaystyle >}{\sim} 10^{23} \displaystyle{\big(\frac{m_\nu}{1\, \mbox{eV}}\big)}$~sec~\cite{EMR}. Furthermore, it is worth mentioning that radiative decays of massive neutrinos with $0.1\stackrel{\displaystyle <}{\sim} m_\nu \stackrel{\displaystyle <}{\sim} 10$~keV have also found some applications in cosmology and astrophysics~\cite{DWS}. In this context, most noticeable is probably the Gunn-Peterson test~\cite{GP}, {\em i.e.}, the search of primordial elements in the intergalactic medium. There seems to be a deficiency of neutral hydrogen and helium in the intergalactic medium~\cite{GP}. A source of photo-ionization of these elements might be a radiatively decaying neutrino. Here, we simply comment on the fact that the mass range of neutrino required for such an explanation is different from what is suggested by the solar neutrino puzzle and the atmospheric neutrino problem. To ionize singly ionized helium, $m_{\nu}$ should be bigger than 109~eV, since the ionization potential is 54.4~eV. A recent investigation of this issue may be found in~\cite{SKS}. \setcounter{equation}{0} \section{Numerical results} \indent In this section, we will present numerical predictions as well as constraints on the basic parameters of our $\not\!\! R$ model, which have been discussed in Section 5. Although there is a large number of parameters that could vary independently, it is important to remark that there exists a strong correlation between new-physics observables and light neutrino masses. This seems to be a generic feature of most of the $R$-parity broken SUSY models considered in the literature~\cite{BBHH,NRV}. However, a novel feature of our minimal $\not\!\! R$ scenario is that the size of the scalar-neutrino vev's and the $\varepsilon_i$ terms can, in principle, be unconstrained. In fact, if $\vec{\Lambda}\simeq 0$ in Eq.~(\ref{31}), which is a form of alignment in the flavour space between the vev's of the sneutrinos, $\vec{w}$, and the $\not\!\! R$ terms, $\vec{\varepsilon}$, this condition alone is sufficient to evade upper limits on the tau-neutrino mass for any value of the SUSY parameters $M$, $M'$, $\mu$, and $\tan\beta$. In order to understand how all new-physics interactions are proportional to $\Lambda_i$ and hence depend on $m_{eff}$ [or $m_{\nu_\tau}$] in Eq.~(\ref{m31}) [Eq.~(\ref{m35})], we evaluate the mixing matrix $\xi$ defined in Eq.~(\ref{xi}). Thus, we have \begin{eqnarray} \xi_{i1} &=& \frac{g'M\mu}{2\, \mbox{det}M_4}\, \Lambda_i\, ,\nonumber\\ \xi_{i2} &=& -\, \frac{g cM\mu}{2\, \mbox{det}M_4}\, \Lambda_i\, ,\nonumber\\ \xi_{i3} &=& \frac{\varepsilon_i}{\mu}\, +\, \frac{(cg^2+g'^2)Mv_2}{4\, \mbox{det}M_4}\, \Lambda_i\, , \nonumber\\ \xi_{i4} &=& -\, \frac{(cg^2+g'^2)Mv_1}{4\, \mbox{det}M_4}\, \Lambda_i \, ,\label{xii} \end{eqnarray} for $i=1\, (e),\, 2\, (\mu)$, and $3\, (\tau)$. It is now easy to see that only the elements $\xi_{i3}$ contain the dominant contributions characterized by being {\em not} proportional to $\Lambda_i$. However, these contributions vanish identically in the relevant expression \begin{equation} \delta_{\nu_i\nu_j}\, -\, \tilde{C}_{\nu_i\nu_j}\ =\ [\tilde{\xi}^* (1-t_z) \tilde{\xi}^T ]_{\nu_i\nu_j}, \end{equation} given in Eq.~(\ref{C}), since the element of the diagonal matrix $(1-t_z)_{33}=0$. Consequently, in the limit of vanishing $\tau$-neutrino mass, the invisible $Z$-boson width predicted in our $\not\!\!\! R$ model will coincide with that found in the SM. Similar strong $m_{\nu_\tau}$ dependence occurs in the non-SM part of the couplings $Zl_il_j$ and $Wl_i\nu_j$ via the mixing matrix $\xi_{{}_L}$, which is given by \begin{eqnarray} (\xi^*_{{}_L})_{i1}= \frac{g\, \Lambda_i}{\sqrt{2} (M\mu-\frac{1}{2}g^2v_1v_2)}\, , \nonumber\\ (\xi^*_{{}_L})_{i2}= \frac{\varepsilon_i}{\mu}\, -\, \frac{g^2v_2\, \Lambda_i}{2\mu (M\mu-\frac{1}{2}g^2v_1v_2)}\, .\label{xiiL} \end{eqnarray} One can readily see that the dominant terms in $\xi_{{}_L}$ are contained in the elements $(\xi_{{}_L})_{i2}$. However, in the $Zl_il_j$ coupling, the new-physics contributions are determined by \begin{equation} [\xi_{{}_L} (1-d) \xi^\dagger_{{}_L}]_{ij}\ =\ (\xi_{{}_L})_{i1} (\xi^*_{{}_L})_{j1}\, , \end{equation} and the elements $(\xi_{{}_L})_{i2}$ always get killed by the diagonal matrix $(1-d)$. Thus, leptonic FCNC $Z$-boson decays and associated universality-breaking effects are proportional to $\Lambda_i$ and are absent if $\nu_\tau$ is massless. Moreover, we find that the non-SM contributions present in the coupling $Wl\nu$ in Eq.~(\ref{BL}) are proportional to \begin{equation} \Big( -\xi_{{}_L}\xi_{{}_L}^\dagger\, -\, \xi^*\xi^T\, +\, 2\xi_{{}_L}t_{{}_L}\xi^T\Big)_{ij}\ =\ -\, (\xi_{{}_L})_{i2}[(\xi^*_{{}_L})_{j2}-\xi_{j3}]\, -\, \xi_{j3}[\xi^*_{i3}\, -\, (\xi_{{}_L})_{i2}]\, .\label{newW} \end{equation} Substituting Eqs.~(\ref{xii}) and (\ref{xiiL}) into Eq.~(\ref{newW}), it is easy to verify that new-physics effects in charged-current interactions are also very strongly correlated with the light neutrino mass $m_\nu$. For reasons mentioned above, we will work in the seesaw approximation by keeping the mass of $\nu_\tau$ finite. For our illustrations, we will consider the following modest $\not\!\! R$ SUSY scenarios: \begin{equation} \begin{array}{r|r|rrrr} \mbox{Scenario}& (\mbox{type of line}) & \tan\beta & M\, \mbox{[GeV]} & \mu\, \mbox{[GeV]} &\varepsilon_\mu\ (\mbox{or}\ \varepsilon_\tau )\, \mbox{[GeV]} \\ \hline \mbox{I} &\mbox{(solid)} & 1 & 50 & 500 & 0 \\ \mbox{II} &\mbox{(dashed)} & 1 & 50 & -50 & -0.5 \\ \mbox{III} &\mbox{(dotted)} & 1 & 100 & 200& 1 \\ \mbox{IV} &\mbox{(dash--dotted)} & 4 & 200 & 400& 2 \\ \end{array} \label{A-scenarios} \end{equation} where $M'=M/2$ and the type of line used in our plots is also indicated. First, we will study possible limits on the $\not\!\! R$ models in Eq.~(\ref{A-scenarios}) that may be derived by the nonobservation of a muon decay into three electrons. In Fig.~1(a), numerical predictions for $B(\mu^-\to e^-e^-e^+)$ as a function of $m_{\nu_\tau}$ are displayed for $w_\mu/w_e=1$. The horizontal dotted line indicates the present experimental limit. Fig.~1(a) also shows the strong quadratic dependence of $B(\mu^-\to e^-e^-e^+)$ on $m_{\nu_\tau}$. In particular, if $w_\mu$ and $w_e$ are comparable in size ({\em e.g.}, $w_\mu/w_e=1$), this constraint is more severe. Qualitatively, we find that \begin{equation} \frac{w_ew_\mu}{w_e^2+w_\mu^2}\, \frac{m_\nu}{M} \ \stackrel{\displaystyle <}{\sim}\ 10^{-6}\, .\label{bound} \end{equation} Of course, this limit gets relaxed for large vev ratios $w_\mu/w_e$. The bound derived from $B(\mu^-\to e^-e^-e^+)$ is more sensitive to the soft-SUSY gaugino mass $M$. To be more precise, our analysis yields the following upper limits on $m_\nu$: \begin{equation} \begin{array}{r|r} \mbox{Scenario}& m_\nu\, \mbox{[MeV]} \\ \hline \mbox{I} & 0.20 \\ \mbox{II} & 0.57 \\ \mbox{III} & 0.43 \\ \mbox{IV} & 0.89 \\ \end{array} \label{A-bound} \end{equation} at 90$\%$ CL. We also remark that $\tau$-lepton number is assumed to be conserved so as to protect a primordial excess in $L_\tau$ from being erased by processes that are in thermal equilibrium. It is then obvious that for scenarios with $w_\mu/w_e=1$ and $M=200$ GeV, $m_{\nu_\tau}<0.9$ MeV. From~(\ref{A-bound}), we see that scenario I gives a stronger limit than the experimental one on the mass of $\nu_\mu$, which is currently $m_{\nu_\mu}<0.27$~MeV at 90$\%$ CL~\cite{PDG94}. As the non-SM couplings depend crucially on the $\tau$-neutrino mass, the less than 1 MeV upper bound on a massive neutrino gives little chance to see new-physics effects in other observables. However, if $\Delta L_e=0$ in the model, {\em i.e.}\ $w_e=\varepsilon_e=0$, inequality~(\ref{bound}) is trivially fulfilled and the so-derived neutrino mass bound does not apply any longer. In Fig.~1(b), numerical estimates reveal that non-SM contributions to the invisible $Z$-boson width are one order of magnitude smaller than the present experimental sensitivity. As a result, experimental searches for physics beyond the SM, based solely on neutrino counting at the $Z$ peak, are bound to be inadequate to unravel the nature of our minimal $\not\!\! R$ model. {}From Fig.~2(a), it can be seen that our minimal $\not\!\! R$ model may predict universality-breaking effects via the observable $U_{br}$ in excess of $10^{-3}$. Such new-physics phenomena might be seen at LEP, if all the experimental data accumulated in the year 1995 are analyzed. Furthermore, in Fig.~2(b), we give theoretical predictions for the observable $R_\pi/R_\pi^{SM}-1$ given in Eq.~(\ref{Rpiexp}). Possible deviations from lepton universality in charged-current interactions turn out to be one order of magnitude smaller than those that can be accessed in experiment. Also, beyond the realm of detection are found to be possible violations of charged-current universality in the decays $\tau\to e \nu\nu$ and $\mu\to e \nu\nu$, which are measured by virtue of the physical quantities $R_{\tau e}$ and $R_{\tau\mu}$. Theoretically, similar is predicted to be the situation for the size of the FCNC $Z$-boson mediated decays, such as $\tau^-\to \mu^- e^- e^+$ and $Z\to ll'$. More explicitly, it is estimated that \begin{eqnarray} B(\tau^-\to \mu^- e^- e^+) &\stackrel{\displaystyle <}{\sim}& 1.\, 10^{-9}\, ,\nonumber\\ B(Z\to l^-l'^+\ \mbox{or}\ l^+l'^-) &\stackrel{\displaystyle <}{\sim}& 1.\, 10^{-8}\, ,\nonumber\\ 1\, -\, \frac{R_{\tau e}}{R_{\tau e}^{SM}} &\stackrel{\displaystyle <}{\sim}& 1.\, 10^{-4}\, ,\nonumber\\ 1\, -\, \frac{R_{\tau\mu}}{R_{\tau\mu}^{SM}} &\stackrel{\displaystyle <}{\sim}& 1.\, 10^{-4}\, . \end{eqnarray} There may also be other places where $R$-parity violation could manifest its presence. Of course, if neutralinos are lighter than the $Z$ boson, one could search for distinctive signatures caused by decays of the form $Z\to \nu_\tau \chi^0$ or $\tau^\pm \chi^\mp$, where $\chi^0$ and $\chi^\pm$ decay subsequently into two $b$-quark jets accompanied by a large amount of missing mass~\cite{BBHH}. However, if the production threshold of heavy neutralinos and charginos is above the LEP centre of mass energy, one then has to rely on studies of possible indirect non-SM signals via sensitive observables devoid of ambiguities coming from the evaluation of hadronic matrix elements, as those discussed in Sections 5.1--5.4. In the same logic, $R$-parity violating effects may also be probed in the $\nu_\mu e$ scattering, even though experimental data do not impose very stringent constraints as compared to those resulting from $B(\mu\to e^-e^-e^+)$~\cite{BGH,BGKLM}. Since our minimal $\not\!\! R$ model only modifies the leptonic sector, one may derive useful constraints from atomic parity violation measurements of the effective `weak charge', $Q_W$, of a heavy nucleus. In the case of $^{133}_{55}$Cs, one has~\cite{BGKLM} \begin{equation} Q_W^{exp}(^{133}_{55}\mbox{Cs})-Q_W^{SM}(^{133}_{55}\mbox{Cs})\ =\ 73.5\cdot [\tilde{\xi}_L (1-d) \tilde{\xi}_L^\dagger]_{11}\ \le\ 3.74 \, ,\label{QW} \end{equation} at 1$\sigma$. The above bound turns out to be rather weak when compared to that derived from $B(\mu\to eee)$. Finally, for reasons that have already been mentioned in Section 5.3, possible limits obtained directly from forward-backward-asymmetry observables similar to $\Delta{\cal A}_{ll'}$ are estimated to be much weaker than those determined by the universality-breaking parameter $U_{br}^{(ll')}$ in Eq.~(\ref{Ubr}), and are therefore not taken into consideration here. \section{Conclusions} \indent The minimal $R$-parity broken SUSY model contains bilinear lepton-number-violating terms ($\varepsilon_i$), which cannot in general be eliminated by a re-definition of the superfields provided soft-SUSY breaking parameters are simultaneously present in the superpotential. The consideration of these $\varepsilon_i$ mass terms, which involve the chiral multiplets of the left-handed leptons and the Higgs field with $Y=+1$, give rise naturally to non-vanishing vev's, $w_i$, of the scalar neutrinos after the spontaneous breaking of the gauge symmetry. In particular, if the vectors $\vec{w}$ and $\vec{\varepsilon}$, spanned in the flavour space, satisfy a kind of alignment relation, $\vec{\Lambda}=0$, forced, {\em e.g.}, by some horizontal symmetry, the afore-mentioned $w_i$ and $\varepsilon_i$ parameters are not restricted by limits on the $\tau$-neutrino mass. Furthermore, constraints from primordial nucleosynthesis and the observed BAU have been considered. Specifically, to evade BAU constraints has been sufficient to impose that at least one separate leptonic number has to be conserved in our $\not\!\! R$ model, {\em e.g.}\ $w_\tau=\varepsilon_\tau=0$ and $w_e=\varepsilon_e=0$. Our main interest has been to investigate the phenomenological implications of this novel $\not\!\!\! R$ model in the light of a number of terrestrial, astrophysical, and cosmological constraints. To be more concrete, we have considered a typical set of $\not\!\! R$ models as is stated in (\ref{A-scenarios}) and confronted it with results obtained from LEP, CLEO and other experiments. We have found that the resulting non-SM contributions to the couplings $Z\nu\nu$, $Zll'$, and $Wl\nu$ show a strong correlation with the $\tau$-neutrino mass and vanish in the massless limit. This direct correlation between the size of $R$-parity-violating phenomena and the magnitude of the neutrino mass appears to be a generic feature of most of the $R$-parity broken models considered in the literature~\cite{BBHH,NRV}. In our analysis, the most severe constraint comes from $B(\mu\to eee)$ for $\not\!\! R$ scenarios, where $\Delta L_\tau=0$, and $L_e$ and $L_\mu$ are maximally violated. In this way, we have been able to set an upper bound on $m_{\nu_\tau}$ by means of Eq.~(\ref{bound}). For instance, for $M=\mu=2M_W$ and $w_e=w_\mu$, we find that $m_{\nu_\tau}\stackrel{\displaystyle <}{\sim} 1$ MeV. Especially, for scenario I in Eq.~(\ref{A-scenarios}), we have $m_\nu < 0.2$~MeV as has been given in Eq.~(\ref{A-bound}), which is even tighter than the current experimental bound on the mass of the $\mu$ neutrino. The remaining observables leave the main bulk of the parameter space unconstrained. The most encouraging prediction is obtained for the universality-violating observable $U_{br}$, with $U_{br}\stackrel{\displaystyle <}{\sim} 2.\, 10^{-3}$. Such phenomena might be seen at LEP, when the analysis of all the data of the year 1995 is completed. For our purposes, we need not study the combined effect of the trilinear $R$-parity-violating couplings $\lambda$ and $\lambda'$, {\em i.e.}\ $\lambda_{ijk}=\lambda'_{ijk}=0$. The reason is that the Yukawa couplings $\lambda_{ijk}$ and $\lambda'_{ijk}$ are not sufficient to explain possible new-physics phenomena that can be shown up in certain low-energy processes and LEP observables, such as $B(l^-\to l'^-l_1^-l^+_1)$, $B(Z\to ll')$, and $U_{br}$, discussed in Sections~5.1--5.3. In this context, we remark that the KARMEN anomaly can, in principle, be explained by assuming the presence of a fourth light neutralino, even though an unnaturally small value of $\mu={\cal O}(10)$~MeV may be required. \vskip1cm \noindent {\bf Acknowledgements.} The authors gratefully acknowledge discussions with Roger Phillips. M.N. would like to thank the theory group of Rutherford Appleton Laboratory for hospitality extended to him during a visit, when part of this work was done. M.N. also gratefully acknowledges financial support by the HCM program under EEC contract no. CHRY-CT 920026. \newpage \setcounter{section}{0} \def\Alph{section}.\arabic{equation}{\Alph{section}.\arabic{equation}} \begin{appendix} \setcounter{equation}{0} \section{Appendix} \indent For completeness, we present that part of the scalar potential (\ref{m9}) that contains the charged singlet fields $E_i$. This also consists of a lepton-number conserving contribution ($V_+^{L}$) and a lepton-number violating one ($V_+^{\not L}$). The former reads \begin{eqnarray} \label{a1} V_+^{L}&=&[\mu^2_{+ij}(E_i^* E_j) + \mbox{H.c.}] + [\mu_{ij}(\phi_2^{\dagger}\varphi_i)E_j + \mbox{H.c.}] +[\mu'_{ij}(\phi_1^{\dagger}\varphi_i)E_j + \mbox{H.c.} ]\nonumber \\ &&+\lambda(\sum_k E_k^* E_k)^2 +(\tilde{\kappa}_{jk} - \lambda \delta_{jk}) (\phi_1^{\dagger}\phi_1)(E_j^* E_k) + \lambda(\phi_2^{\dagger}\phi_2)(E_k^{*} E_k) \nonumber \\ &&+ (\mu_{ijnm} - \lambda \delta_{in}\delta_{jm})(\varphi_i^{\dagger} \varphi_n)(E_j^{*} E_m) + 4\tilde{\kappa}_{nmij}(\varphi_n^{\dagger} \varphi_i)(E_m^{*} E_j)\, , \end{eqnarray} where we have defined \begin{eqnarray} \label{a2} && \lambda = {1 \over 2} g'^2\, , \nonumber \\ && \mu_{ij} = \mu^* h_{ij}\, , \nonumber \\ && \tilde{\kappa}_{jk}= h_{ij}^* h_{ik}=\tilde{\kappa}_{kj}^*\, , \nonumber \\ && \mu_{ijnm}= h_{ij}^* h_{nm}= \mu_{nmij}^* \, ,\nonumber \\ && \tilde{\kappa}_{nmij}=\lambda_{knm}^* \lambda_{kij}= \tilde{\kappa}_{ijnm}^*\, . \end{eqnarray} In Eq.~(\ref{a2}), $h_{ij}$ and $\lambda_{ijk}$ are couplings from the superpotential (\ref{m3}) and (\ref{m4}), respectively. $\mu_{+ij}^2$ and $\mu'_{ij}$ are soft-SUSY breaking parameters from Eq.~(\ref{m8}). For the lepton-number-violating contribution, we obtain \begin{equation} \label{a3} V_+^{\not L}= -2i\tilde{\kappa}_{ijk}(\phi_1^T \tau_2 \varphi_j)(E_i^* E_k) + \kappa'_{ijk}(\varphi_i^T \tau_2 \varphi_j)E_k + \mbox{H.c.}\, , \end{equation} where \begin{equation} \label{a4} \tilde{\kappa}_{ijk} = h_{ni}^* \lambda_{njk}\, , \end{equation} and $\kappa'_{ijk}$ is a soft-SUSY breaking parameter contained in Eq.~(\ref{m8}). By analogy with Eq.~(\ref{m20}), from $V_+^L + V_+^{\not L}$, we can derive conditions for not having CP violation in this part of the potential. These conditions are listed below \begin{eqnarray} \label{a5} && \mu^*_{ijnm}\eta^*_{{}_{L_n}} \eta_{{}_{L_i}} \eta^*_{{}_{+m}} \eta_{{}_{+j}}= \mu_{ijnm}\, , \nonumber \\ && \tilde{\kappa}^*_{jk} \eta^*_{{}_{+k}} \eta_{{}_{+j}}=\tilde{\kappa}_{jk} \, ,\nonumber \\ &&\mu^*_{ij}\eta_{{}_{2}} \eta^*_{{}_{L_i}} \eta^*_{{}_{+j}}=\mu_{ij} \, ,\nonumber \\ &&\mu'^{*}_{ij}\eta_{{}_{1}} \eta^*_{{}_{L_i}} \eta^*_{{}_{+j}}=\mu'_{ij} \, ,\nonumber \\ &&\tilde{\kappa}^*_{nmij}\eta^*_{{}_{L_n}} \eta_{{}_{L_i}} \eta^*_{{}_{+m}} \eta_{{}_{+j}}= \tilde{\kappa}_{nmij}\, , \nonumber \\ &&\tilde{\kappa}^*_{ijk}\eta^*_{{}_{1}} \eta^*_{{}_{L_j}} \eta_{{}_{+i}} \eta^*_{{}_{+k}}= \tilde{\kappa}_{ijk}\, , \nonumber \\ && \kappa'^{*}_{ijk}\eta^*_{{}_{L_i}} \eta^*_{{}_{L_j}} \eta^*_{{}_{+k}} =-\kappa'_{ijk}\, , \end{eqnarray} where summation convention is not implied. In Eq.~(\ref{a5}), $\eta_{{}_{+j}}$ are the CP phases of the scalar fields $E_j$, similar to the notation of Eq.~(\ref{m20}). In general, both sets, (\ref{m20}) and (\ref{a5}), should not be viewed independently of one another. For instance, using the equalities in Eqs.~(\ref{m20}) and (\ref{a5}), one can derive (no summation convention) \begin{eqnarray} \label{a6} && \Im m(\lambda_6 \mu_{ij}\mu'^{*}_{nm}\tilde{\kappa}_{mj}\kappa_{in})=0 \, ,\nonumber \\ && \Im m(\lambda_6 \mu^*_{ijnm} \mu_{nm}\mu'^{*}_{ij})=0 \, ,\nonumber \\ && \Im m(\mu^*_{ijnm}\kappa_{ni}\tilde{\kappa}^*_{mj})=0\, , \end{eqnarray} and many similar relations of this kind. \end{appendix} \newpage
\section{Introduction} \label{introduction} \cite{sandt:presupdrt} introduces \cite{lewis}' notion of {\em accommodation\/} in Discourse Representation Theory \cite{KR:DtoL} as a tool to account for gaps in the discourse. His theory of presupposition projection takes presuppositions to behave like anaphora. Anaphoric expressions normally are linked to antecedents that have previously been established in the discourse. If example~\ref{kingoffrance} would appear in a context where no {\sl king of France} is present -- hence no antecedent is available -- then Van der Sandt's algorithm {\em accommodates\/} the existence of a king of France. \begin{examples} \ex\exlab{kingoffrance} When I give a party, the king of France always attends it. \end{examples} This is different from the situation where a definite description can be {\em linked\/} to an antecedent that was previously introduced by an indefinite description, as in~\ref{celebrity}. There is no need to accommodate an antecedent, because there is already a suitable candidate available. \begin{examples} \ex\exlab{celebrity} When I invite a celebrity, the celebrity never comes. \end{examples} Example~\ref{barkeeper} however is slightly different. There is no actual antecedent for the anaphoric expression {\sl the barkeeper}, but because of {\sl a bar}, there isn't really a problem, apparently there is some implicit antecedent. Van der Sandt's projection algorithm fails to make this implicit link, and accommodates the existence of {\sl a barkeeper} to the global context, in fact no theory on presupposition that we know of can deal with these\footnote{ The closest comes probably Beaver's dynamic theory of presupposition \cite{beaver}. }. \begin{examples} \ex\exlab{barkeeper} When I go to a bar, the barkeeper always throws me out. \end{examples} Contrasting~\ref{barkeeper} with~\ref{playground} makes our point even clearer; this sentence sounds truly infelicitous. The hearer tries to somehow link this {\sl barkeeper} with familiar information, and fails. \begin{examples} \ex[?]\exlab{playground} When I go to a playground, the barkeeper always throws me out. \end{examples} {\sl A bar} provides sufficient information to license {\sl the barkeeper}, but in {\sl a playground} there is nothing that can establish such a link. Making a link between the new discourse referent (i.e. {\sl the barkeeper}) to the network of discourse referents that is already established, is called {\em bridging\/} (\cite{bridging}, \cite{heim:diss}). Definite descriptions that can be bridged to existing information do not need the accommodation of new referents; example~\ref{kingoffrance} requires accommodation, but \ref{barkeeper} can be solved with mere bridging. An adequate theory of presupposition obviously needs a serious explanation of bridging to account for the projection problem of presupposition.\\ To account for these phenomena, we borrow from \cite{pus:GL} and compare {\em bridging\/} with {\em coercion\/}. Pustejovsky presents examples like~\ref{beginabook}: \begin{examples} \ex\exlab{beginabook} I would like to begin a new book tonight. \end{examples} Here, too, some information is missing: {\sl begin} implies some event, but {\sl a new book} is an artifact. The fact that the speaker should be interpreted as {\sl beginning to read the book}, or -- if he is a writer -- {\sl to write} one, is motivated by what we know about {\sl book}. Pustejovsky claims that such information should be considered lexical knowledge of the noun, which is represented in a so-called {\em qualia structure\/}. Based on this information, arguments of improper types can be {\em coerced\/} to proper ones. We will see how a similar approach can be followed to account for {\em bridging\/}. In section~\ref{qscoercion} we will present Pustejovsky's ideas in more detail, explaining concepts like {\em coercion\/} and {\em qualia structure\/}. In Pustejovsky's work these ideas only get applied on the sentence level. Section~\ref{bridging} will show how the ideas of Van der Sandt and Pustejovsky fit together very nicely, even complementing each other and we will show that {\em bridging\/} operates intra- as well as inter-sentential. In section~\ref{examples} we will present some examples of {\em linking\/}, {\em bridging\/} and {\em accommodation\/}, and in section~\ref{functionalcomposition} we will discuss the notion of functional composition and coercion in this model. \section{Qualia Structure and Coercion} \label{qscoercion} \subsection{Qualia Structure} \label{qscoercionqs} In \cite{pus:GL} and subsequent papers the notions of {\em coercion} and {\em qualia structure} have been introduced. Qualia structure can be seen as a set of lexical entailments. For instance, the word {\sl book} entails at least the two events of {\sl reading} and {\sl writing} it, besides the knowledge that it consists of several separate parts, like the {\sl cover}, {\sl pages}, etc. Pustejovsky suggests four {\em qualia roles} to represent such knowledge: {\sc formal}, {\sc constitutive}, {\sc telic} and {\sc agentive}. In \cite{pus:GL} these have been defined as follows\footnote{ In more common AI-related terms we could rephrase them as: {\sc formal} - {\sc isa}, {\sc constitutive} - {\sc part-of~/~hasa}, {\sc telic} - {\sc purpose} and {\sc agentive} - {\sc cause}.}: \begin{itemize} \item {\sc formal}: That which distinguishes the object within a larger domain. \item {\sc constitutive}: The relation between an object and its constituents or proper parts. \item {\sc telic}: Purpose and function of the object. \item {\sc agentive}: Factors involved in the origin or "bringing about" of an object. \end{itemize} \label{important} The exact structure of this kind of lexical semantic knowledge seems to be very intricate. Again for the same example, it is important for instance to realize that a {\sl book} is at the same time a {\em physical object} and an {\em information container}. The first description considers the physical viewpoint, whereas the second defines the conceptual angle of what constitutes our idea of a {\sl book}. What angle one takes ({\em physical object} or {\em information container}) has immediate consequences for the knowledge that is represented in the rest of the qualia structure. The composing parts of the physical side of a book ({\sl pages}, {\sl cover}, etc.) are different from those of the conceptual side ({\sl title}, {\sl sections}, {\sl paragraphs}, etc.). The same goes for the representation of typical events a book is involved in. The physical 'quality' (qual) of a book can be {\sl printed}, {\sl typeset} or even {\sl shelved}. The information 'quality' can be said to undergo the events of {\sl reading}, {\sl writing} as mentioned before. It is however undeniable that the two main qualities of book, along with all their entailments, are intimately related to each other and should be represented accordingly in one comprehensive (qualia) structure. \subsection{Coercion} The need for a rich lexical semantic knowledge representation like qualia structure becomes clear in considering sentences like \ref{beginabook} above, which is repeated here: \begin{examples} \ex\exlab{beginabook2} I would like to begin a new book tonight. \end{examples} As mentioned before, the verb {\sl begin} expects an event here but has to settle for an {\sl artifact} (book). We can now use the qualia structure of this {\sl artifact} to infer some {\sl event} that is entailed by it and which can stand in its place. This is an example of what Pustejovsky has called {\em metonymic reconstruction} \cite{pus:GL} for cases where an interpretation can be inferred from some partial meaning of the word in question. In more general terms, anytime a word or phrase is not of the desired type\footnote{ Possibly this use of the term {\em type} is not appropriate and we should use {\em sort} instead. However, here we present the terms as they have been defined by \cite{pus:GL}.} (like {\sl artifact}, {\sl event}, etc.) we are allowed to {\em coerce} it into one of its entailments that is of the appropriate type, where the entailments are stored in its qualia structure. Another example of this is the following sentence: \begin{examples} \ex\exlab{announcenewmodel} BMW announced a new model. \end{examples} Here the verb {\sl announce} is looking for a subject of type {\em animate} while only one of type {\em institute} is available. The qualia structure of any {\em institute} however should represent the fact that they are made up of people, which are {\em animate} entities. So, in this sentence we can infer that some human at the BMW company did the actual announcement.\\ This summarizes Pustejovsky's program as described in \cite{pus:GL} and subsequent papers. In this paper we extend coercion with the notion of context, which seems not only a valid research topic but also desperately needed because of the restricted explaining power of coercion if context is not considered. Take for instance sentence~\ref{beganabook} : \begin{examples} \ex\exlab{beganabook} John began a book. \end{examples} Although above we assumed several times that one can infer {\sl read} and {\sl write} events from the qualia structure of {\sl book} in order to make this sentence semantically well formed, this can only be a default approximation. We would need an actual context for this sentence to decide what event exactly should be inferred. Imagine for instance a dinner, organized by the {\sc literary and culinary society}, where all dishes are shaped in the form of books$\ldots$\footnote{ Still another problem concerning the lack of context is illustrated by the following examples where no argument at all is available for coercion to take place, {\sl Monday} and {\sl yesterday} are modifiers: \begin{quote} I propose Monday.\\ I began yesterday. \end{quote} Pustejovsky (personal communication) has termed this loosely as {\em null coercion}, because although coercion should take place it cannot be executed properly. Taking context into account could be of help however to make the sentences sound more natural, as the following examples show: \begin{quote} Let's make an appointment. I propose Monday.\\ Let's play darts. I begin. \end{quote} It seems that null coercion should coerce an anaphor which is of the required type. In both examples this would be event-type anaphors. } This example is farfetched, but it may make the point more clear. We do not assume that the qualia structure of {\sl book} should contain any reference to this particular example. It is important however to realize that any {\sl artifact} entails by default a number of events in which it is engaged. In this particular context these events would be overruled. \section{Bridging in DRT with Qualia Structure} \label{bridging} This section shows how we deal with anaphora resolution in general, and particularly bridging, in a version of DRT which uses extensively {\em qualia\/} information. We define the language of Discourse Representation Structures (DRSs) of our extended DRT, show how resolution works, and finally give some detailed examples. \subsection{A Sketch of the Architecture} Basically, we extend Van der Sandt's theory of presupposition with the notion of bridging anaphora. In short, Van der Sandt views presupposition as anaphora with more descriptive content, and uses one and the same mechanism for dealing with both phenomena \cite{sandt:presupdrt}. Anaphoric information can either be resolved to an antecedent that is available from discourse, or if no antecedent is found, be accommodated. We add a possibility of bridging to the resolution algorithm. The basic architecture of the system is: \begin{enumerate} \item parse sentence: result is a sentence-DRS \item merge sentence-DRS with main-DRS \item perform anaphora resolution \end{enumerate} A {\em sentence-DRS} is a DRS with all anaphoric information unresolved, and is the result of a bottom-up driven semantic construction dependent on some syntactic structure. A sentence-DRS can be viewed as a sort of under-specified logical form with respect to anaphoric information. Special types of DRSs ($\alpha$-DRSs) mark anaphoric information. The {\em main-DRS} is the DRS of the context interpreted so far. It is a {\em proper} DRS, i.e., a DRS with no unresolved anaphoric information. Proper DRSs can be interpreted as in standard DRT: they are {\em true} with respect to a certain model if they can be {\em embedded\/} in that model (\cite{KR:DtoL}. Before we explain how anaphora resolution works we define DRSs and the merging operation. \subsection{Discourse Representation Structures} Let's introduce some terminology. Discourse markers are variables ranging over objects in the domain. Terms are either discourse markers or DRSs. Furthermore, we adopt a typed lambda-calculus for DRSs \cite{tilburg,muskens:cdrt}. DRSs are defined as follows: \begin{quote} {\bf Definition 1. DRS}\\ If $U$ is a set of discourse markers, $C$ is a set of DRS-Conditions, and t$_1$,...,t$_n$ terms, then $<U,C>$ is a DRS, $<U,C> \oplus <U^\prime,C^\prime>$ is a DRS, $\lambda$ t$_1$,...,t$_n$. $<U,C>$ is a DRS. Nothing else is a DRS. \end{quote} \begin{quote} {\bf Definition 2. DRS-Conditions}\\ If x$_1$,..,x$_n$ are discourse markers, P an n-place condition, K and K$_1$ DRSs, then P(x$_1$,...,x$_n$), x$_1$ = x$_2$, K $\to$ K$_1$, $\lnot$ K, K $\lor$ K$_1$, $\alpha$:K, and Q:K are DRS-Conditions. Nothing else is a DRS-Condition. \end{quote} The first five DRS-Conditions we already know from standard DRT \cite{KR:DtoL} and need no further explanation. So called $\alpha$-DRSs represent unresolved anaphoric information. DRSs that contain $\alpha$-DRSs are therefore unresolved DRSs. Q-DRSs represent qualia structure, with Q$_F$ for {\sl formal}, Q$_C$ for constitutive, Q$_A$ for agentive and Q$_T$ for telic. For notational purposes we use $\cal{Q}$ to represents a {\em set\/} of qualia-DRSs\footnote{ As mentioned in section~\ref{qscoercionqs} on page~\pageref{important}, the distinction that is made in the formal role carries through in all other qualia roles. This could be represented by embedding Q$_C$, Q$_A$ and Q$_T$ in Q$_F$. This is beyond the scope of our paper.}. Now for merging: \begin{quote} {\bf Definition 3. Merging ($\oplus$)}.\\ $<U_1,C_1> \oplus <U_2,C_2> = <U_1 \cup U_2, C_1 \cup C_2>$ \end{quote} The merge operation takes two DRSs and makes a union of the sets of discourse markers and a union of the sets conditions. Merging of DRSs is used both for constructing DRSs (cf. \cite{tilburg}) and {\em coercive accommodation}. The latter term brings us to the next definition. Qualia-information, represented in Q-DRSs is normally not accessible and does not affect the truth-conditions of a DRS. It is introduced in the lexicon and brought into discourse via the DRS bottom-up construction algorithm. If necessary, for example to play the role of antecedent, the qualia structure is put forward to the surface by a process we call {\em coercive accommodation}. It is defined as a function from DRSs to sets of DRSs: \begin{quote} {\bf Definition 4. Coercive Accommodation (CA).}\\ {\sc ca}($<U,C>$) = $\{ <U,C> \oplus$ K $|$ Q:K $\in C\}$ \end{quote} Note that CA is always local: it cannot accommodate qualia information which is embedded. Note also that we have defined CA only for DRS without lambda's: this will do for the purposes of this paper. Q-DRSs are also used for {\em type coercion}, which is discussed later on in this paper. In DRT the structure of DRSs restricts the choice of possible antecedents of an anaphoric construction. For a discourse marker to be the antecedent for an anaphor, it must be {\em accessible} from the DRS which the anaphor is represented. To define accessibility of DRSs and discourse markers we first use the notion {\em subordination} between DRSs. We adopt the notation C(K) meaning the set of conditions of DRS (K), and U(K) meaning the set of discourse markers of K. \begin{quote} {\bf Definition 5. Subordination.}\\ If K$_1$, K$_2$, and K$_3$ are DRSs, then K$_2$ is subordinated to K$_1$ (or K$_1$ subordinates K$_2$) if K$_1$ $\oplus$ K$_2$, K$_1$ $\to$ K$_2$ $\in$ C(K$_3$), K$_2$ $\to$ K$_3$ $\in$ C(K$_1$), $\lnot$ K$_2$ $\in$ C(K$_1$), K$_2$ $\lor$ K$_3$ $\in$ C(K$_1$), K$_3$ $\lor$ K$_2$ $\in$ C(K$_1$), $\alpha$:K$_2$ $\in$ C(K$_1$), Q:K$_2$ $\in$ C(K$_1$), and K$_2$ is subordinated to K$_3$ and K$_3$ is subordinated to K$_1$. \end{quote} So, if x is a discourse marker and K$_1$ and K$_2$ are DRSs, and x is in the domain of K$_1$ (x $\in$ U(K$_1$)), then x is accessible from K$_2$ if K$_2$ is subordinated to K$_1$. \subsection{Anaphora Resolution} Left to explain is how anaphora resolution works. We repeat for convenience that resolution can take place in three different ways: \begin{enumerate} \item resolution to an accessible, suitable discourse marker ({\em linking}) \item resolution to coercively accommodated material of an accessible DRS ({\em bridging}) \item accommodation of the anaphoric information to an accessible DRS ({\em accommodation}) \end{enumerate} We introduced accessibility already, but haven't explained yet the notion of `suitable' discourse marker, or better: suitable DRSs. Suitability is an extra constraint on the choice of antecedent. A DRS is suitable to another DRS if there is a way you find a match between discourse markers and conditions between both. More formally: \begin{quote} {\bf Definition 6. Suitability.}\\ A DRS K$_2$ is $m$-suitable to DRS K$_1$ if there is a mapping $m$ such that scope($m$)=U(K$_2$) and for every x it is the case that m(x) $\in$ U(K$_1$) and there is a DRS K$_3$ such that C(K$_3$) $\subseteq$ C(K$_1$) if U(K$_3$) = $\{$ m(x) $|$ x $\in$ U(K$_1$)$\}$. \end{quote} We now introduce the heart of the system: anaphora resolution. This algorithm works as follows. All anaphoric information in the main-DRS (of course after merging it with the sentence-DRS of the last processed sentence) is resolved. This information is clearly marked because these are just our $\alpha$-DRSs. Resolution either unifies this material with a suitable antecedent or accommodates it, and as a result, $\alpha$-DRSs disappear. After resolving all $\alpha$-DRSs, we are left with a proper-DRS, a DRS which is fully specified with respect to anaphoric information. This DRS is model-theoretically interpretable, as in standard DRT. To describe the component, we use K$_\alpha$ to indicate anaphoric DRSs, and K$_m$ for the main-DRS. Definition 7 describes a function that takes a certain main-DRS and a certain $\alpha$-DRS from it, and returns a set of DRSs (since there could be more than one possible antecedent or accommodation site) with this $\alpha$-DRS resolved. The output of this function can be fed back into the same function until all anaphoric information is resolved (all $\alpha$-DRS have been consumed).\footnote{The order of which resolution of anaphoric structure takes place is important as well. We don't pay any attention to this, but see \cite{sandt:presupdrt}.} \begin{quote} {\bf Definition 7. Anaphora Resolution.}\\ \begin{tabular}{ll} {\sc ar}(K$_\alpha$,K$_m$) & = $\{$ K' $|$ K' $\in$ {\sc link}(K$_\alpha$,K$_m$) $\}$ iff $|${\sc link}(K$_\alpha$,K$_m$)$|$ $>$ 0\\ & = $\{$ K' $|$ K' $\in$ {\sc bridge}(K$_\alpha$,K$_m$) $\}$ iff \parbox[t]{44mm}{$|${\sc link}(K$_\alpha$,K$_m$)$|$ $=$ 0 and\\ $|${\sc bridge}(K$_\alpha$,K$_m$)$|$ $>$ 0}\\ & = $\{$ K' $|$ K' $\in$ {\sc acc}(K$_\alpha$,K$_m$) $\}$ iff \parbox[t]{44mm}{$|${\sc link}(K$_\alpha$,K$_m$)$|$ $=$ 0 and\\ $|${\sc bridge}(K$_\alpha$,K$_m$)$|$ $=$ 0}\\ \end{tabular} \end{quote} Note that this definition prefers linking to bridging, and bridging to accommodation, which we assume is right. {\sc link}, {\sc bridge}, and {\sc acc} are functions from the main DRS to sets of DRSs. We use DRSsubstitution to describe these operations ($[$ K$_1$ / K$_2$ $]$ K$_3$ means that K$_1$ is substituted for K$_2$ in K$_3$). \begin{quote} {\bf Definition 8. Linking.}\\ {\sc link}(K$_\alpha$,K$_m$) = $\{$ \parbox[t]{100mm}{ $[$ K$_3$ / K$_2$ $]$ K$_m$ $|$ K$_\alpha$ is subordinated and m-suitable to K$_1$ \& $\alpha$:K$_\alpha$ $\in$ C(K$_2$) \& U(K$_3$)=U(K$_2$) $\cup$ U(K$_\alpha$) \& C(K$_3$)=C(K$_2$)-$\alpha$:K$\alpha$ $\cup$ C(K$_\alpha$) $\cup$ $\{$ x=y $|$ m(x)=y $\}$ $\}$ } \end{quote} \begin{quote} {\bf Definition 9. Bridging.}\\ {\sc bridge}(K$_\alpha$,K$_m$) = $\{$ \parbox[t]{100mm}{ K'$_m$~$|$~K$_\alpha$ is subordinated K$_4$ \& K$_1$ $\in$ {\sc ca}(K$_4$) \& \linebreak m-suitable to K$_1$ \& $\alpha$:K$_\alpha$ $\in$ C(K$_2$) \& \linebreak U(K$_3$)=U(K$_2$) $\cup$ U(K$_\alpha$) \& C(K$_3$)=C(K$_2$)-$\alpha$:K$\alpha$ $\cup$ C(K$_\alpha$) $\cup$ $\{$~x=y~$|$~m(x)=y~$\}$ \& K'$_m$ = $[$ K$_3$ / K$_2$ $]$ K$_m$ \& K'$_m$ = $[$~K$_1$~/~K$_4$~$]$~K$_m$ $\}$ } \end{quote} \begin{quote} {\bf Definition 10. Accommodation.}\\ {\sc acc}(K$_\alpha$,K$_m$) = $\{$ \parbox[t]{100mm}{ K'$_m$ $|$ K$_\alpha$ is subordinated to K$_1$ \& $\alpha$:K$_\alpha$ $\in$ C(K$_2$) \& U(K$_3$)=U(K$_2$) \& C(K$_3$)=C(K$_2$)-$\alpha$:K$_\alpha$ \& K'$_m$=$[$~K$_1$~$\oplus$~K$_\alpha$~/~K$_1$ $]$ K$_m$ \& K'$_m$=$[$~K$_3$ / K$_2$ $]$ K$_m$~$\}$ } \end{quote} Accommodation has its limits. First, it shouldn't introduce free variables, and Van der Sandt introduces a number of acceptability rules for accommodation. These are briefly: resolution should not introduce contradictions and require a contribution to discourse. For more discussion on this issue the interested reader should consult \cite{sandt:presupdrt}. \subsection{Examples of Linking, Bridging and Accommodation} \label{examples} This section exemplifies the notions {\em linking}, {\em bridging}, and {\em accommodation}, which we introduced in the previous section. We will do this in view of the examples given in the introduction. For each of these examples we give the DRS with all anaphoric information unresolved, and the fully resolved derived after anaphora resolution as well. For reasons of clarity, only the relevant parts of the DRSs are deeply analyzed. \subsubsection{Linking} The first example involves simple linking between anaphor and antecedent. Consider the unresolved DRS of \ref{celebrity}: \begin{center} \drs{}{\\ \drs{x}{celebrity(x)\\ I-invite(x)} $\to$ \drs{}{\\$\alpha$:\drs{y}{celebrity(y)}\\ never-comes(y)}\\} \end{center} The definite description introduces an $\alpha$-DRS for {\sl the celebrity}, since this is presupposed information. Trying to link this anaphoric information is successful, since there is an accessible suitable discourse marker available. The result is the resolved DRS: \begin{center} \drs{}{\\ \drs{x}{celebrity(x)\\ I-invite(x)} $\to$ \drs{y}{celebrity(y)\\ y=x\\ never-comes(y)}\\} \end{center} This DRS can be read as: {\sl If I invite a celebrity, he never comes}. \subsubsection{Bridging} Now for our bridging\footnote{ Bridging does not seem to be the preferred option in the case of resolution of pronouns as the following examples show: \begin{quote} When I go to a bar, he always throws me out.\\ When BMW announced a new model, he looked very proud. \end{quote} In both sentences a reading for 'he' can be found by linking to coercively accommodated material out of the Q-DRS from respectively {\sl bar} and {\sl BMW}, i.e. {\sl a barkeeper} or {\sl a spokesperson}. However they don't seem to be the preferred readings as has been shown by \cite{McGlashan,SanfordandGarrod} for similar examples.} example. The unresolved DRS of example~\ref{barkeeper} is (simplifying the Q-DRS for convenience): \begin{center} \drs{}{\\ \drs{x}{bar(x)\\ Q:\drs{z}{barkeeper(z)\\of(z,x)}\\ I-go-to(x)} $\to$ \drs{}{\\ $\alpha$:\drs{y}{barkeeper(y)}\\ always-throws-me-out(y)}\\} \end{center} The presupposition trigger {\sl the barkeeper} introduces the anaphoric information. Linking fails, the only available discourse marker is not suitable since the condition of anaphoric information does not match with it. Bridging is successful, though, yielding the resolved DRS: \begin{center} \drs{}{\\ \drs{x z}{bar(x)\\ $\cal{Q}$\\ I-go-to(x)\\ barkeeper(z)\\ of(z,x)} $\to$ \drs{y}{barkeeper(y)\\ y=z\\ always-throws-me-out(y)}\\} \end{center} This DRS does not assume a particular barkeeper that throws the speaker out, but a barkeeper that belongs to the bar the speaker goes to -- the correct prediction. \subsubsection{Accommodation} Accommodation is our emergency case: {\em if everything fails, then accommodate}. This happens in cases like \ref{kingoffrance}, which unresolved DRS is: \begin{center} \drs{}{\\ \drs{x}{party(x)\\ I-give(x)} $\to$ \drs{}{\\ $\alpha$:\drs{y}{king-of-france(y)}\\ \\ $\alpha$:\drs{z}{}\\ always-attends(y,z)}\\} \end{center} The pronoun represented by the discourse marker {\sl z} can be linked to {\sl x}. But we cannot link {\sl the king of France} to some accessible discourse marker, nor is there a way to make bridging inference. The only possibility left is to accommodate the king: \begin{center} \drs{y}{king-of-france(y)\\ \drs{x}{party(x)\\ I-give(x)} $\to$ \drs{z}{z=x\\ always-attends(y,z)}\\} \end{center} This DRS represents the reading: {\sl there is a king of France, and if I give a party, he will attend it}. This is again the correct prediction. \section{Functional Composition and Coercion} \label{functionalcomposition} \subsection{Defining the notions} Functional Composition, including type coercion, is defined as follows, K$_1$ being the functor, K$_2$ the argument, and $\vec{\sigma}$ a sequence of terms such that K$_2$(t)($\vec{\sigma}$) is a proposition: \begin{quote} {\bf Definition 11. Functional Composition ($\odot$).}\\ \begin{tabular}{ll} K$_1$ $\odot$ K$_2$ & = $\lambda$ $\vec{\sigma}$. K$_1$($\lambda$ v. (K$_2$(v)($\vec\sigma$))) iff K$_1$ is of type $<\alpha$,t$>$ and v is of type $\alpha$; \\ & = K$_1$ $\odot$ K$_3$ (where K$_3$ $\in$ {\sc tc}(K$_2$)) otherwise. \end{tabular} \end{quote} Clause one is like the functional composition rule (in \cite{tilburg}). This rule has the nice property that it doesn't need type-shifting of arguments. It always binds the first argument position of the argument, and has functional application as a special case ($\vec{\sigma}$ is empty then). The second clause does the type coercion stuff (cf. \cite{pus:TC}): \begin{quote} {\bf Definition 12. Type Coercion (TC).}\\ {\sc tc}(K)=$\{$ K$^\prime$ $\odot$ K $|$ K$^\prime$ $\in$ {\sc qa}(K) $\}$ \end{quote} \begin{quote} {\bf Definition 13. Qualia Access (QA)}.\\ {\sc qa}(K)=$\{$ K$_Q$ $|$ Q:K$_Q$ $\in$ C(K) or {\sc qa}(K') where K' is a sub-DRS of K $\}$ \end{quote} Note that {\sc tc} also works for arbitrarily deep embedded DRSs by use of the Qualia Access function. This is nice for quantified NPs like {\sl every book}, where the qualia DRSs lexically introduced for {\sl book} has been placed in the restrictor. \subsection{Some lexical entries} In this section we present some example lexical entries. In this paper we will only assign (a simplified) qualia structure to nouns\footnote{ \cite{pus:book} assumes qualia structures for {\em all} categories.}, see {\bf book}. Lexical entries can be abbreviated by their boldface notation -- {\bf write} stands for the semantic part of the lexical entry of {\sl write}. We use small e,x,y and z for variables over type e (for entities, i.e. objects and events), capital P for DRS of type $<$e,t$>$ (properties), and capital E for event-types (normally $<e,t>$). \begin{tabbing} {\bf pres} \= : \= \kill {\bf book}\>:\> $\lambda$z.\drs{z}{book(z)\\ \qdrs{F}{}{info\_cont(z)}\\ \\ \qdrs{C}{Z} {sections(Z)\\ has(z,Z)}\\ \\ Q$_{\mbox{\tiny A}}$: {\bf write} \\ Q$_{\mbox{\tiny T}}$: {\bf read}} \end{tabbing} By introducing determiners ({\bf a, the, every}) we account for the possibility to carry qualia structure through the derivation. Note the difference between these three determiners. The article {\em the} introduces an $\alpha$-DRS since it is a presupposition trigger. \begin{tabbing} {\bf pres} \= : \= \kill {\bf a} \>:\> $\lambda$ P$_1$ P$_2$. \drs{x}{} $\oplus$ P$_1$(x) $\oplus$ P$_2$(x)\\ \\ {\bf the} \>:\> $\lambda$ P$_1$ P$_2$. \drs{}{\mbox{}\\[-2ex] $\alpha$: \drs{x}{} $\oplus$ P$_1$(x)\mbox{}\\[-2ex]} $\oplus$ P$_2$(x)\\ \\ {\bf every} \>:\> $\lambda$ P$_1$ P$_2$. \drs{}{\mbox{}\\[-2ex]\drs{x}{} $\oplus$ P$_1$(x) $\to$ P$_2$(x)\mbox{}\\[-2ex]} \end{tabbing} The proper name {\bf john} introduces an anaphoric DRS which is merged with the representation of its predicate. Proper names do {\em not\/} have qualia structure (see footnote earlier). \begin{tabbing} {\bf pres} \= : \= \kill {\bf john}\>:\> $\lambda$ P. \drs{}{\mbox{}\\[-2ex]$\alpha$:\drs{x}{john(x)}\mbox{}\\[-2ex]} $\oplus$ P(x) \end{tabbing} The verbs {\bf write} and {\bf read} introduce event-types. Lambda-operators bind the variables that will fulfill the thematic roles {\em agent\/} and {\em theme}. \begin{tabbing} {\bf pres} \= : \= \kill {\bf write} \>:\> $\lambda$ y x e. \drs{}{write(e)\\ agent(e,x)\\ theme(e,y)}\\ \\ {\bf read} \>:\> $\lambda$ y x e. \drs{}{read(e)\\ agent(e,x)\\ theme(e,y)} \end{tabbing} The aspectual verb {\bf begin} expects something that expresses an event-type. We here simply treat it as a modifier, and ignore its further aspectual presuppositions. Finally, tense {\bf pres} applies to an event-type and binds off the event variable:the result is a DRS of type t, i.e. a DRS with no lambda variables. \begin{tabbing} {\bf pres} \= : \= \kill {\bf begin} \>: \> $\lambda$ E x e. \drs{}{begin(e)} $\oplus$ E(x)(e) \\ \\ {\bf pres} \> : \> $\lambda$ E. \drs{e}{now(e)} $\oplus$ E(e) \end{tabbing} \subsection{A sample derivation} Let us now follow the derivation of `John begins a book'. Functional composition of {\bf a} with {\bf book} yields a noun phrase that contains the qualia structure of the noun and awaits a property to merge with. \begin{quote} {\bf a} $\odot$ {\bf book} = $\lambda$P.\drs{z}{book(z)\\ \qdrs{F}{}{info\_cont(z)}\\ \\ \qdrs{C}{Z} {sections(Z)\\ has(z,Z)}\\ \\ Q$_{\mbox{\tiny A}}$: {\bf write} \\ Q$_{\mbox{\tiny T}}$: {\bf read}} $\oplus$ P(z) \end{quote} Functional composition of {\bf begin} with {\bf a $\odot$ book} can only work with a type coercion. The event that {\bf begin} requires cannot be found directly, so no simple link can be made. From the qualia structure of {\bf a $\odot$ book} we can for example coerce {\bf read}, and this qualifies as the required event. This coercion step is worked out later. \begin{quote} {\bf begin} $\odot$ ({\bf a} $\odot$ {\bf book}) = $\lambda$ x. e. \drs{y} {begin(e)\\ read(e)\\ agent(e,x)\\ theme(e,y)\\ book(y)\\ $\cal{Q}$} \end{quote} The rest of the derivation follows straightforwardly. {\bf begin $\odot$ (a $\odot$ book)} functionally composed with {\bf john} results in a lambda-DRS. \begin{quote} {\bf john} $\odot$ ({\bf begin} $\odot$ ({\bf a} $\odot$ {\bf book})) = $\lambda$ e. \drs{y} {\mbox{}\\[-2ex]$\alpha$:\drs{x}{john(x)\mbox{}\\[-2ex]}\\ begin(e)\\ read(e)\\ agent(e,x)\\ theme(e,y)\\ book(y)\\ $\cal{Q}$ } \end{quote} Adding tense ({\bf pres}) to the lambda-DRS turns it into a proper DRS -- all anaphoric information has been resolved and no antecedent for the presupposed event needed to be accommodated. \begin{quote} {\bf pres} $\odot$ ({\bf john} $\odot$ ({\bf begin} $\odot$ ({\bf a} $\odot$ {\bf book}))) = \drs{e y}{\mbox{}\\[-2ex]$\alpha$:\drs{x}{john(x)\mbox{}\\[-2ex]}\\ now(e)\\ begin(e)\\ read(e)\\ agent(e,x)\\ theme(e,y)\\ book(y)\\ $\cal{Q}$} \end{quote} \pagebreak Naturally we could just as easily have taken {\bf write} instead of {\bf read}, or for that matter, any of the other events that occur in the qualia structure of {\bf book}. Since {\bf read} and {\bf write} are the only events, the result of coercing {\bf a book} is as follows:\\ \begin{minipage}{\textwidth} {\sc tc}($\lambda$P.\drs{z}{book(z)\\ \qdrs{F}{}{info\_cnt(z)}\\ \\ \qdrs{C}{Z} {sections(Z)\\ has(z,Z)}\\ \\ Q$_{\mbox{\tiny A}}$: {\bf write} \\ Q$_{\mbox{\tiny T}}$: {\bf read}} $\oplus$ P(z)) = $\{$ $\lambda$x.e.\drs{z} {book(z)\\ write(e)\\ agent(e,x)\\ theme(e,z)\\ $\cal{Q}$}, $\lambda$x.e.\drs{z}{book(z)\\ read(e)\\ agent(e,x)\\ theme(e,z)\\ $\cal{Q}$}$\}$ \end{minipage} \section{Conclusions and Further Work} \label{conclusions} We have shown that Bridging and Coercion can be seen in very much the same light, viz. as using implicit lexical information to accommodate a missing antecedent. In doing so, we have extended Pustejovsky's ideas on Coercion and placed it in a discourse perspective. On the other hand we have extended Van der Sandt's algorithm with Bridging, and thus made it more complete with respect to the linguistic data. The work presented here is limited to definite descriptions; we have not looked into other presupposition triggers. \cite{beaver} mentions the following examples, where inferencing takes place. \begin{examples} \ex\exlab{beaver1} Probably, if Jane takes a bath, Bill will be annoyed that there is no more hot water. \ex\exlab{beaver2} If Spaceman Spiff lands on Planet X, he will be bothered by the fact that his weight is higher than it would be on Earth. \end{examples} In \ref{beaver1} the inference is made that taking a bath uses up a hot water reservoir, in \ref{beaver2} that landing on a strange planet may make changes to your weight. To fit with these examples in the framework we presented in this paper remains for future research.
\section{Introduction} The formation of cosmological structure in the universe, inhomogeneities in the matter distribution like quasars at redshifts up to $z\sim 5$, galaxies, clusters, super clusters, voids and walls, is an outstanding basically unsolved problem within the standard model of cosmology. At first sight it seems obvious that small density enhancements can grow sufficiently rapidly by gravitational instability. But global expansion of the universe and radiation pressure counteract gravity, so that, e.g., in the case of a radiation dominated, expanding universe no density inhomogeneities can grow faster than logarithmically. Even in a universe dominated by pressure-less matter, cosmic dust, the growth of density perturbations is strongly reduced by the expansion of the universe. On the other hand, we know that the universe was extremely homogeneous and isotropic at early times. This follows from the isotropy of the 3K Cosmic Microwave Background (CMB), which represents a relic of the plasma of baryons, electrons and radiation at times before protons and electrons combined to hydrogen. After a long series of upper bounds, measurements with the COsmic Background Explorer satellite (COBE) have finally established anisotropies in this radiation \cite{Sm} at the level of \[ \langle{\Delta T\over T}(\theta)\rangle \sim 10^{-5} ~~~\mbox{ on angular scales}~ 7^o\le \theta\le 90^o ~.\] On smaller angular scales the observational situation is at present somewhat confusing and contradictory \cite{WJNPW}, but many upper limits require $ \Delta T/T < 4\times 10^{-5} $ on all scales $\theta <8^o$. All observations together clearly rule out the simplest model of a purely baryonic universe with density parameter $\Omega \sim 0.1$ and adiabatic initial fluctuations (either the initial perturbations are too large to satisfy CMB limits, or they are too small to develop into the observed large scale structure). The most conservative way out, where one just allows for non--adiabatic initial perturbations (minimal isocurvature model), also faces severe difficulties \cite{PS,GSS,PD,HBS}. In other models one assumes that initial fluctuations are created during an inflationary epoch, but that the matter content of the universe is dominated by hot or cold dark matter or a mixture of both. Dark matter particles do not interact with photons other than gravitationally and thus induce perturbations in the CMB only via gravitation. In these models, inflation generically leads to $\Omega =1$, while the baryonic density parameter is only $\Omega_Bh^2 \sim 0.02$, compatible with nucleosynthesis constraints. With one component of dark matter, these models do not seem to agree with observations \cite{GSS,Os}, however, if a suitable mixture of hot and cold dark matter is adopted, the results from numerical simulations look quite promising \cite{Ho,XK,DSS}, although they might have difficulties to account for the existence of clusters at a redshift $z\sim 1$ \cite{Ba}. In these dark matter models initial fluctuations are generated during an inflationary phase. Since all worked out models of inflation face difficulties (all of them have to invoke fine tuning to obtain the correct amplitude of density inhomogeneities), we consider it very important to investigate yet another possibility: Density perturbations in the dark matter and baryons might have been triggered by seeds. Seeds are an inhomogeneously distributed form of energy which makes up only a small fraction of the total energy density of the universe. Particularly natural seeds are topological defects. They can form during symmetry breaking phase transitions in the early universe \cite{Ki,CDTY}. Depending on the symmetry being gauged or global, the corresponding defects are called local or global. The fluctuation spectrum on large scales observed by COBE is not very far from scale invariant \cite{Go}. This has been considered a great success for inflationary models which generically predict a scale invariant fluctuation spectrum. However, as we shall see, also models in which perturbations are seeded by global topological defects yield scale invariant spectra of CMB fluctuations. To be specific, we shall mainly consider texture, $\pi_3$--defects which lead to event singularities in four dimensional spacetime \cite{Tu,d95}. Global defects are viable candidates for structure formation, since the scalar field energy density, $\rho_S$, of global topological defects scales like $\rho_S \propto 1/(at)^{2}$ (up to a logarithmic correction for global strings) and thus always represents the same fraction of the total energy density of the universe ($t$ is conformal time). \begin{equation} \rho_S/\rho \sim 8\pi G\eta^2 \equiv 2\epsilon ~, \end{equation} where $\eta$ determines the symmetry breaking scale ({\sl see} Fig.~1). For the background spacetime we assume a Friedmann--Lema\^{\i}tre universe with $\Omega=1$ dominated by cold dark matter (CDM). We choose conformal coordinates such that \[ ds^2 = a^2(-dt^2 + \delta_{ij}dx^idx^j) ~. \] Numerical analysis of CMB fluctuations from topological defects on large scales has been performed in \cite{BR,PST}; a spherically symmetric approximation is discussed in \cite{DHZ}. Results for intermediate scales angular are presented in \cite{CFGT}. All these investigations (except \cite{DHZ}) use linear cosmological perturbation theory in synchronous gauge and (except \cite{PST}) take into account only scalar perturbations. Here we derive a fully gauge invariant and local system of perturbation equations. The (non--local) split into scalar, vector and tensor modes on hyper surfaces of constant time is not performed. We solve the equations numerically in a cold dark matter (CDM) universe with global texture. In this paper, we detail the results outlined in a previous letter \cite{DZ}. Furthermore, we present explicit derivations of the equations, a description of our numerical methods and we briefly discuss some tests of our codes. Since there are no spurious gauge modes in our initial conditions, there is no danger that these may grow in time and some of the difficulties to choose correct initial conditions ({\sl see} e.g. \cite{PST}) are removed. However, as we shall discuss in Section~3, the results do depend very sensitively on the choice of initial conditions. Nevertheless, we should keep in mind, that we are investigating models of structure formation which rely on the particle physics and cosmology at temperatures of $T\sim T_{GUT} \sim 10^{16}GeV$. An energy scale about which we have no experimental evidence whatsoever. The physical model adopted for our calculations should thus always be considered as a toy model, of which we hope it captures the features relevant for structure formation of the 'realistic physics' at these energies. Therefore, we suggest, not to take the results serious much beyond about a factor of two or so. On the other hand, our models show that the particle physics at GUT scale may have left its traces in the distribution of matter and radiation in the present universe, yielding the exciting possibility to learn about the physics at the highest energies, smallest scales, by probing the largest structures of the universe. \vspace{0.2cm} We calculate the CMB anisotropies on angular scales which are larger than the angle subtended by the horizon scale at decoupling of matter and radiation, $\theta> \theta_d$. For $\Omega =1$ and $z_d \approx 1000$ \begin{equation} \theta_d = 1/\sqrt{z_d+1} \approx 0.03 \approx 2^o ~. \end{equation} It is therefore sufficient to study the generation and evolution of microwave background fluctuations after recombination. During this period, photons stream freely, influenced only by cosmic gravitational redshift and by perturbations in the gravitational field (if the medium is not re-ionized). In Section~2 we derive a local and gauge invariant perturbation equation to calculate the CMB fluctuations. In Section~3, we put together the full system of equations which has to be solved to investigate gravitationally induced CMB fluctuations and the dark matter perturbation spectrum in a model with global topological defects. We discuss the choice of initial conditions and the numerical treatment of this system in Section~4. The next section is devoted to the presentation and analysis of our numerical results. We end with conclusions in Section~6. \vspace{0.2cm} {\bf Notation:} We denote conformal time by $t$. Greek indices run from 0 to 3, Latin indices run from 1 to 3. The metric signature is chosen $(- + + +)$. We set $\hbar=c=k_{Boltzmann}=1$ throughout. \section{A Local and Gauge Invariant Treatment of the Perturbed Liouville Equation} Collision-less particles are described by their one particle distribution function which lives on the seven dimensional phase space \[ {\cal P}_m = \{ (x,p)\in T\!{\cal M}| g(x)(p,p)=-m^2\} ~. \] Here $\cal M$ denotes the spacetime manifold and $T{\cal M}$ its tangent space. The fact that collision-less particles move on geodesics translates to the Liouville equation for the one particle distribution function, $f$. The Liouville equation reads \cite{Ste} \begin{equation} X_g(f)=0 \label{L}~.\end{equation} In a tetrad basis $(e_\mu)_{\mu=0}^3$ of $\cal M$, the vector field $X_g$ on ${\cal P}_m$ is given by ({\sl see} e.g. \cite{Ste}) \begin{equation} X_g = (p^\mu e_\mu - \omega^i_{\:\mu}(p)p^\mu{\partial \over \partial p^i}) ~, \label{liou} \end{equation} where $\omega^\nu_{\:\mu}$ are the connection 1--forms of $({\cal M},g)$ in the basis $e^\mu$, and we have chosen the basis \[(e_\mu)_{\mu=0}^3~~ \mbox{ and }~~~ ({\partial\over \partial p^i})_{i=1}^3 ~~\mbox{ on }~~~~ T{\cal P}_m~, ~~~~~~ p=p^\mu e_\mu~. \] We now apply this general framework to the case of a perturbed Friedmann universe. The metric of a perturbed Friedmann universe with density parameter $\Omega=1$ is given by $ds^2=g_{\mu\nu} dx^\mu dx^\nu$ with \begin{equation} g_{\mu\nu} = a^2(\eta_{\mu\nu} + h_{\mu\nu}) = a^2 \tilde{g}_{\mu\nu} ~, \end{equation} where $(\eta_{\mu\nu}) = diag(-,+,+,+)$ is the flat Minkowski metric and $(h_{\mu\nu})$ is a small perturbation, $|h_{\mu\nu}|\ll1$. We now use the fact that the motion of photons is conformally invariant: We show that for massless particles and conformally related metrics, \[g_{\mu\nu}= a^2\tilde{g}_{\mu\nu}~,\] \begin{equation} (X_gf)(x,p)=0 ~~~\mbox{ is equivalent to }~~~ (X_{\tilde{g}}f)(x,ap)=0 ~. \label{conform} \end{equation} This is easily seen if we write $X_g$ in a coordinate basis: \[ X_g = b^\mu\partial_\mu -\Gamma_{\alpha\beta}^ib^\alpha b^\beta{\partial\over \partial b^i} ~,\] with \[ \Gamma_{\alpha\beta}^i={1\over 2}g^{i\mu}(g_{\alpha\mu},_\beta + g_{\beta\mu},_\alpha-g_{\alpha\beta},_\mu) ~. \] The $b^\mu$ are the components of the momentum $p$ with respect to the {\em coordinate} basis: \[ p=p^\mu e_\mu = b^\mu\partial_\mu ~.\] If $(e_\mu)$ is a tetrad with respect to $g$, then $\tilde{e}_\mu=ae_\mu$ is a tetrad basis for $\tilde{g}$. Therefore, the coordinates of of $ap=ap^\mu\tilde{e}_\mu= a^2p^\mu e_\mu = a^2b^\mu\partial_\mu$ with respect to $\partial_\mu$ on $({\cal M},\tilde{g})$ are given by $a^2b^\mu$. In the coordinate basis thus our statement Eq.~(\ref{conform}) follows, if we can show that \begin{equation} (X_{\tilde{g}}f)(x^\mu,a^2b^i)=0 ~~~\mbox{ iff }~~~ (X_gf)(x^\mu,b^i)=0 \label{conformco} \end{equation} Setting $v=ap=v^\mu\tilde{e}_\mu = w^\mu\partial_\mu$, we have $v^\mu=ap^\mu$ and $w^\mu=a^2b^\mu$. Using $p^2=0$, we obtain the following relation for the Christoffel symbols of $g$ and $\tilde{g}$: \[ \Gamma^i_{\alpha\beta}b^\alpha b^\beta= \tilde{\Gamma}^i_{\alpha\beta}b^\alpha b^\beta +{2a,_\alpha\over a}b^\alpha b^i ~.\] For this step it is crucial that the particles are massless! For massive particles the statement is of course not true. Inserting this result into the Liouville equation we find \begin{equation} a^2X_gf = w^\mu(\partial_\mu f|_b -2{a,_\mu\over a} b^i{\partial f\over \partial b^i}) -\tilde{\Gamma}_{\alpha\beta}^iw^\alpha w^\beta {\partial f\over \partial w^i} ~, \label{lstar} \end{equation} where $\partial_\mu f|_b$ denotes the derivative of $f$ w.r.t. $x^\mu$ at constant $(b^i)$. Using \[\partial_\mu f|_b = \partial_\mu f|_w + 2{a,_\mu\over a}b^i{\partial f\over \partial b^i} ~,\] we see, that the braces in Eq.~(\ref{lstar}) just correspond to $\partial_\mu f|_w$. Therefore, \[a^2X_g f(x,p)=w^\mu\partial_\mu f|_w-\tilde{\Gamma}_{\alpha\beta}^iw^\alpha w^\beta {\partial f\over \partial w^i} = X_{\tilde{g}}f(x,ap) ~. \] We have thus shown that the Liouville equation in a perturbed Friedmann universe is equivalent to the Liouville equation in perturbed Minkowski space, \begin{equation} (X_{\tilde{g}}f)(x,v)=0 ~, \label{LM} \end{equation} with $v=v^\mu\tilde{e}_\mu = ap^\mu\tilde{e}_\mu$.\footnote{Note that also Friedmann universes with non vanishing spatial curvature, $K\neq 0$, are conformally flat and thus this procedure can also be applied for $K\neq 0$. Of course, in this case the conformal factor $a^2$ is no longer just the scale factor but depends on position. A coordinate transformation which transforms the metric of $K\neq 0$ Friedmann universes into a conformally flat form can be found, e.g., in \cite{CDD}.} We now want to derive a perturbation equation for Eq.~(\ref{LM}). If $\bar{e}^\mu$ is a tetrad in Minkowski space, $\tilde{e}_\mu = \bar{e}_\mu + {1\over 2}h_\mu^\nu\bar{e}_\nu$ is a tetrad w.r.t the perturbed geometry $\tilde{g}$. For $(x,v^\mu\bar{e}_\mu )\in \bar{P}_0$, thus, $(x,v^\mu\tilde{e}_\mu)\in \tilde{P}_0$. Here $\bar{P}_0$ denotes the zero mass one particle phase space in Minkowski space and $\tilde{P}_0$ is the phase space with respect to $\tilde{g}$, perturbed Minkowski space. We define the perturbation of the distribution function $F$ by \begin{equation} f(x,v^\mu \tilde{e}_\mu) = \bar{f}(x,v^\mu \bar{e}_\mu) + F(x,v^\mu\bar{e}_\mu) ~. \end{equation} Liouville's equation for $f$ then leads to a perturbation equation for $F$. We choose the natural tetrad \[\tilde{e}_\mu=\partial_\mu -{1\over 2}h_\mu^\nu\partial_\nu\] with the corresponding basis of 1--forms \[\tilde{\theta}^\mu=dx^\mu +{1\over 2}h^\mu_\nu dx^\nu ~.\] Inserting this into the first structure equation, $d\tilde{\theta}^\mu= -\omega^\mu_{~~\nu}\wedge dx^\nu$, one finds \[ \omega_{\mu\nu}=-{1\over 2}(h_{\mu\lambda},_\nu - h_{\nu\lambda},_\mu)\theta^\lambda ~.\] Using the background Liouville equation, namely that $\bar{f}$ is only a function of $v=ap$, we obtain the perturbation equation \[ (\partial_t +\gamma^i\partial_i)F = -{v\over 2}[(\dot{h}_{i0}-h_{00},_i)\gamma^i +(\dot{h}_{ij}-h_{0j},_i)\gamma^i\gamma^j]{d\bar{f}\over dv} ~,\] where we have set $v^i=v\gamma^i$, with $v^2=\sum_{i=1}^3(v^i)^2$. Let us parameterize the perturbations of the metric by \begin{equation} \left(h_{\mu\nu}\right) = \left(\begin{array}{ll} -2A & B_i \\ B_i & 2H_L\delta_{ij} +2H_{ij} \end{array}\right), \label{scalar} \end{equation} with $H_i^i=0$. Inserting this above we obtain \begin{equation} (\partial_t +\gamma^i\partial_i)F = -[\dot{H}_L +(A,_i +{1\over 2}\dot{B}_i)\gamma^i + (\dot{H}_{ij}-{1\over 2}B_{i,j})\gamma^i\gamma^j]v{d\bar{f}\over dv} ~. \label {LF} \end{equation} From Eq.~(\ref{LF}) we see that the perturbation in the distribution function in each spectral band is proportional to $v{d\bar{f}\over dv}$. This shows ones more that gravity is achromatic. We thus do not loose any information if we integrate this equation over photon energies. We define \[ m = {\pi\over \rho_ra^4}\int Fv^3dv ~.\] $4m$ is the fractional perturbation of the brightness $\iota$, \[ \iota = a^{-4} \int f v^3dv ~. \] This is obtained using the relation \begin{equation} {4\pi}\int {d\bar{f}\over dv}v^4dv = -4\int \bar{f}v^3dvd\Omega =-4\rho_ra^4 ~. \label{rel} \end{equation} Setting $\iota = \bar{\iota}(T(\gamma,x))$, one finds that $ \iota =(\pi/60) T^4(\bm{\gamma},x)$. Hence, $m$ corresponds to the fractional perturbation in the temperature, \begin{equation} T(\gamma,x) = \bar{T}(1+m(\gamma,x)) ~.\label{T} \end{equation} Another derivation of Eq.~(\ref{T}) is given in \cite{d94}. Since the $v$ dependence of $F$ is of the form $v{d\bar{f}\over dv}$, we have with Eq.~(\ref{rel}) \[ F(x^\mu,\gamma^i,v)=-m(x^\mu,\gamma^i)v{d\bar{f}\over dv} ~. \] This shows that $m$ is indeed the quantity which is measured in a CMB anisotropy experiment, where the spectral information is used to verify that the spectrum of perturbations is the derivative of a blackbody spectrum. Of course, in a real experiment located at a fixed position in the Universe, the monopole and dipole contributions to $m$ cannot be measured. They cannot be distinguished from a background component and from a dipole due to our peculiar motion w.r.t. the CMB radiation. Multiplying Eq.~(\ref{LF}) with $v^3$ and integrating over $v$, we obtain the equation of motion for $m$ \begin{equation} \partial_tm+\gamma^i\partial_im= \dot{H}_L +(A,_i +{1\over 2}\dot{B}_i)\gamma^i + (\dot{H}_{ij}-{1\over 2}B_i,_j)\gamma^i\gamma^j ~. \label{Lm} \end{equation} It is well known that the equation of motion for photons only couples to the Weyl part of the curvature (null geodesics are conformally invariant). The r.h.s. of Eq.~(\ref{Lm}) is given by first derivatives of the metric only which could at most represent integrals of the Weyl tensor. To obtain a local, non integral equation, we thus rewrite Eq.~(\ref{Lm}) in terms of $\triangle m$. It turns out, that the most suitable variable is however not $\triangle m$ but $\chi$, which is given by \[ \chi = \triangle m - (\triangle H_L-{1\over 2}H,_{ij}^{ij}) - {1\over 2}(\triangle B_i-3\partial^j\sigma_{ij})\gamma^i ~, \] \[ \mbox{where }~~ \sigma_{ij}= -{1\over 2}(B_i,_j+B_j,_i) + {1\over 3}\delta_{ij}B_l^{,l} +\dot{H}_{ij}. \] Note that $\chi$ and $\triangle m$ only differ by the monopole contribution, $\triangle H_L-(1/2)H^{ij},_{ij}$ and the dipole contribution, $(1/2)(\triangle B_i -3\partial^j\sigma_{ij})\gamma^i$. The higher multipoles of $\chi$ and $\triangle m$ agree. An observer at fixed position and time cannot distinguish a monopole contribution from an isotropic background and a dipole contribution from a peculiar motion. Only the higher multipoles, $l\ge 2$ contain information about temperature anisotropies. For a fixed observer therefore, we can identify $\triangle^{-1}\chi$ with $\delta T/T$. In terms of metric perturbations, the electric and magnetic part of the Weyl tensor are given by ({\sl see}, e.g. \cite{Ma,d94}) \begin{eqnarray} E_{ij} &=& {1\over 2}[\triangle_{ij}(A-H_L) -\dot{\sigma}_{ij} -\triangle H_{ij}-{2\over 3}H_{lm}^{,lm}\delta_{ij} + H_{il}^{,l},_j + H_{jl}^{,l},_i] \label{E} \\ B_{ij} &=& -{1\over 2}(\epsilon_{ilm}\sigma_{jm},_l + \epsilon_{jlm}\sigma_{im},_l ) ~, \label{B} \end{eqnarray} \[ \mbox { with }~~ \triangle_{ij} =\partial_i\partial_j -(1/3)\delta_{ij}\triangle ~.\] Explicitly working out $(\partial_t+\gamma^i\partial_i)\chi$ using Eq.~(\ref{Lm}), yields after some algebra the equation of motion for $\chi$: \begin{equation} (\partial_t +\gamma^i\partial_i)\chi = 3\gamma^i\partial^jE_{ij} +\gamma^k\gamma^j\epsilon_{kli} \partial_lB_{ij} \equiv S_T(t,\bm{x},\bm{\gamma})~ , \label{Lchi} \end{equation} where $\epsilon_{kli}$ is the totally antisymmetric tensor in three dimensions with $\epsilon_{123}=1$. The spatial indices in this equation are raised and lowered with $\delta_{ij}$ and thus index positions are irrelevant. Double indices are summed over irrespective of their positions. In eqn. Eq.~(\ref{Lchi}) the contribution from the electric part of the Weyl tensor does not contain tensor perturbations. On the other hand, scalar perturbations do not induce a magnetic gravitational field. The second contribution to the source term in Eq.~(\ref{Lchi}) thus represents a combination of vector and tensor perturbations. If vector perturbations are negligible, the two terms on the r.h.s of Eq.~(\ref{Lchi}) yield a split into scalar and tensor perturbations which is local. Since the Weyl tensor of Friedmann Lema\^{\i}tre universes vanishes, the r.h.s. of Eq.~(\ref{Lchi}) is manifestly gauge invariant (this is the so called Stewart--Walker lemma \cite{SW}). Hence also the variable $\chi$ is gauge invariant. Another proof of the gauge invariance of $\chi$, discussing the behavior of $F$ under infinitesimal coordinate transformations is presented in \cite{d94}. The general solution to Eq.~(\ref{Lchi}) is given by \begin{equation} \chi(t,\bm{x},\bm{\gamma}) = \int_{t_i}^t S_T(t',\bm{x}+(t'-t)\bm{\gamma}, \bm{\gamma})dt' ~ + ~ \chi(t_i,\bm{x}+(t_i-t)\bm{\gamma}, \bm{\gamma}) ~, \label{chi} \end{equation} where $S_T$ is the source term on the r.h.s. of Eq.~(\ref{Lchi}). Let us compare this result with the more familiar one, where one calculates $\delta T/T$ by integrating photon geodesics (which is of course equivalent to solving the Liouville equation). For simplicity, we specialize to the case of pure scalar perturbations (the expressions for vector and tensor perturbations given in \cite{d94} can be compared with Eq.~(\ref{chi}) in the same manner.) For scalar perturbations, integration of photon geodesics yields \cite{d94} \begin{equation} {\delta T\over T}(t_f,\bm{x}_f,\bm{n}) = -[{1\over 4}D_g^{(r)} + V_i\cdot n^i + (\Psi-\Phi)]^f_i + \int_i^f(\dot{\Psi}-\dot{\Phi})d\lambda ~. \label{geo} \end{equation} Here $\Psi$ and $\Phi$ denote the Bardeen potentials as defined, e.g., in \cite{KS,d94}. On super horizon scales (which are the important scales for the Sachs--Wolfe) contribution $V_i\cdot n^i$ can be neglected. Furthermore, the contributions in the square bracket of Eq.~(\ref{geo}) from the final time $t=t_f$, only lead to uninteresting monopole and dipole terms. We now use that the electric contribution to the Weyl tensor for purely scalar perturbations is given by \cite{d94}) \[ E_{ij} = {1\over 2}(\partial_i\partial_j -{1\over 3}\triangle)(\Psi-\Phi) \equiv {1\over 2}\triangle_{ij}(\Psi-\Phi) ~.\] Therefore $\partial_i(\Psi-\Phi)=3\partial^jE_{ij}$. Using furthermore \[ -(\Psi-\Phi)\left|^f_i \right. = -\int_i^f[\dot{\Psi}-\dot{\Phi}+ (\Psi-\Phi),_in^i]d\lambda ~, \] Eq.~(\ref{geo}) leads to \begin{equation} {\delta T\over T}(t,\bm{x},\bm{n}) = {1\over 4}D_g^{(r)}(t_i,\bm{x}_i) - 3\int_i^f\triangle^{-1}\partial^jE_{ij}n^i dt ~ \label{geo2} \end{equation} If we take into account that the direction $\bm{n}$ in Eq.~(\ref{geo}), the direction of an {\sl incoming } photon corresponds to $-\bm{\gamma}$ in Eq.~(\ref{chi}), we find that Eq.~(\ref{geo}) coincides with Eq.~(\ref{chi}) for scalar perturbations, and that \begin{equation} \chi(t_i,\bm{x}_i,\bm{\gamma})= {1\over 4}\triangle D_g^{(r)}(t_i,\bm{x}_i) ={1\over 4}\triangle D_g^{(r)}(t_i,\bm{x} -\bm{\gamma}(t-t_i)) ~. \end{equation} We now want to investigate this initial value and decompose Eq.~(\ref{geo2}) into terms due to CDM and terms coming from the source, the scalar field. We assume that dark matter and radiation perturbations are adiabatic on {\em superhorizon scales}, \[D_g^{(r)}=(4/3)D_g^{(c)}~.\] Since radiation and CDM probably have been a single fluid at early times (e.g. at the time of the phase transition), this assumption is reasonable. It is however, inconsistent to set $D_g^{(r)} =4/3D_g^{(c)}$ on subhorizon scales. Due to the different equations of state for the two components, adiabaticity cannot be maintained on sub-horizon scales \cite{KS}. We can then derive from the equations given in \cite{d94} \[{1\over 4}D_g^{(r)} ={5\over 3}\Phi_C + {2\over 3}\dot{\Phi}_C/(\dot{a}/a) +\Phi^S ~.\] Here the Bardeen potentials are split into parts due to cold dark matter ($_C$) and the scalar field ($_S$) respectively. For cold dark matter $\Psi_C=-\Phi_C$. Using this, we can bring Eq.~(\ref{chi}) into the form \begin{eqnarray} {\delta T\over T}(t_f,\bm{x}_f,\bm{n}) &=& {1\over 3}\Psi_C(t_i,\bm{x}_i) - {2\over 3}\dot{\Psi}_C/(\dot{a}/a)(t_i,\bm{x}_i) +2\int_i^f\dot{\Psi}_C dt \nonumber \\ && +\Phi^S(t_i,\bm{x}_i) - 3\int_i^f\triangle^{-1}S_{TS}(t,\bm{x}_f-(t_f-t)\bm{n},\bm{n}) dt ~, \label{genu} \end{eqnarray} where $S_{TS}$ denote the portion of the source term due to the scalar field only: \begin{equation} S_{TS}= -3n^i\partial^jE^{(S)}_{ij} +n^kn^j\epsilon_{klj} \partial_lB^{(S)}_{ij} ~. \label{STS} \end{equation} Eq. (\ref{genu}) is much better suited for numerical investigation than the general expression Eq.~(\ref{chi}). This can be demonstrated by considering the case of pure CDM without source term: In this case $\Phi_C=-\Psi_C=$ constant and from Eq.~(\ref{genu}) we easily recover the well--kown result \[ {\delta T\over T}(t,\bm{x},\bm{n}) = {1\over 3}\Psi_C(t_i,\bm{x}-\bm{n}(t-t_i)) ~,\] whereas Eq.~(\ref{chi}) in this case leads to \[ {\delta T\over T}(t,\bm{x},\bm{n}) = {\delta T\over T}(t_i,\bm{x}_i,\bm{n}) + 2\Psi_C(t_i,\bm{x}_i) ~.\] In other words, the unknown initial condition in Eq.~(\ref{chi}) cancels $5/6$ of the naive result for the case of adiabatic CDM fluctuations. Even though due to the existence of $\dot{\Psi}_C$ terms. the cancelation is slightly less substantial in our case, the assumption of adiabaticity on superhorizon scales is a crucial ingredient of the model. The electric and magnetic parts of the Weyl tensor are determined by the perturbations in the energy momentum tensor via Einstein's equations. We assume that the source for the geometric perturbations is given by the scalar field and dark matter. The contributions from radiation may be neglected. Furthermore, vector perturbations of dark matter (which decay quickly) are neglected. The divergence of $E_{ij}$ is then determined by ({\sl see} Appendix~A) \begin{equation} 3\partial^jE_{ij} = 8\pi G\rho_{C}a^2D_i + 8\pi G(\partial_i\delta T_{00} +3({\dot{a}\over a})\delta T_{0i} - (3/2)\partial^j\tau_{ij}) \label{dE} ~, \end{equation} where the first term on the r.h.s. is the dark matter source term, $\rho_{C}$ denoting the dark matter energy density. The second contribution is due to the scalar field: The energy momentum tensor of the scalar field \[ T_{\mu\nu}^{S} = \phi,_\mu\phi,_\nu - {1\over 2}g_{\mu\nu}\phi^{,\lambda}\phi,_\lambda \] yields \[\tau_{ij} \equiv T_{ij} -(a^2/3)\delta_{ij}T^l_l = \tau_{ij}^{S} = \phi,_i\phi,_j -(1/3)\delta_{ij}(\nabla\phi)^2 ~,\] \[ \delta T_{0j}= T_{0j}^{S} = \dot{\phi}\phi,_j ~~, \] \[ \delta T_{00}= T_{00}^{S} = {1\over 2}((\dot{\phi})^2 + (\nabla\phi)^2) ~,\] and $D_j$ is a gauge invariant perturbation variable for the density gradient. For scalar perturbations $ D_j = \partial_jD$. The evolution equation for the dark matter density perturbation is given by ({\sl see} \cite{d90}) \begin{equation} \ddot{D} + ({\dot{a}\over a})\dot{D} - 4\pi Ga^2\rho_{C}D = 8\pi G \dot{\phi}^2 \label{C} ~. \end{equation} During the radiation dominated era $8\pi G\rho_RD_R$ in principle has to be included in Eq.~(\ref{C}). But since radiation perturbations quickly decay on sub-horizon scales, and since dark matter fluctuations cannot grow in a radiation dominated universe \cite{Me}, their influence is not relevant. The equation of motion for $B_{ij}$ is more involved. A somewhat cumbersome derivation ({\sl see Appendix~A}) yields \begin{equation} a^{-1}(aB_{ij})^{\cdot\cd} -\triangle B_{ij} = 8\pi G{\cal S}^{(B)}_{ij} ~, \label{Bij} \end{equation} \[ \mbox{with }~~~ {\cal S}^{(B)}_{ij}= -\epsilon_{lm(i}\delta T_{0l},_{j)m} +\epsilon_{lm(i}\dot{\tau}_{j)l},_m ~.\] Here $(i...j)$ denotes symmetrization in the indices $i$ and $j$. To the source term ${\cal S}^{(B)}$ only vector and tensor perturbations contribute. It is thus entirely determined by the energy momentum tensor of the scalar field. \section{The System of equations for Global Scalar Field Induced Fluctuations} In this section we collect all the equations which determine the system under consideration. We also repeat equations which have already been derived in Section~2. Let us begin with the scalar field equation of motion. The energy momentum tensor of the scalar field is a small perturbation. In first order perturbation theory, we can thus solve the equation of motion of the scalar field in the background, Friedmann--Lema\^{\i}tre geometry, neglecting geometric perturbations. The equation of motion for the scalar field $\phi$ is then given by \begin{equation} g^{\mu\nu}\nabla_\mu\nabla_\nu\phi + {\partial V\over \partial\phi} = 0 ~, \label{phi1}\end{equation} where $g^{\mu\nu}$ denotes the unperturbed metric and $\nabla_\mu$ is the covariant derivative with respect to this metric. For our numerical computations, we consider an $O(4)$ model. In $O(N)$ models the scalar field, $\phi\in \bm{ R}^N$ and the zero temperature potential is given by $V_0={\lambda\over 4}(\phi^2-\eta^2)^2$ for some energy scale $\eta$. At high temperatures, $T>T_c \sim \eta$, one loop corrections to the effective potential dominate and the minimum of the effective potential is at $\phi=0$. Below the critical temperature the minimum is shifted (in the simplest case) to $<\phi^2> =(1-(T/T_c)^2)\eta^2$ ({\sl see} \cite{Ki,d95} and references therein). The vacuum manifold, i.e. the space of minima of the effective potential, then becomes a $(N-1)$--sphere, $\bm{ S}^{(N-1)}$. Since \[ \pi_k(\bm{ S}^m) = \left\{ \begin{array}{ll} 0~~, & k<m\\ \bm{ Z}~~, k=m, \end{array}\right. \] the lowest non--vanishing homotopy group of a $m$--sphere is always $\pi_m$. Since probably higher defects are unstable and decay into lower ones\footnote{This is an unproven conjecture, motivated, e.g., by observations of the density of textures and monopoles in liquid crystals and by numerical experiments \cite{CDTY,Kinew}}, the $m$--sphere is a suitable vacuum manifold to study $\pi_m$ defects. If the system under consideration is at a temperature $T$ much below the critical temperature, $T\ll T_c$, it becomes more and more improbable for the field $\phi$ to leave the vacuum manifold. $\phi$ will leave the vacuum manifold only if it would otherwise be forced to gradients of order $ (\nabla \phi)^2 \sim \lambda\phi^2\eta^2$, thus only over length scales of order $l=1/(\sqrt{\lambda}\eta) \equiv m_\phi^{-1}$ ($l$ is the transversal extension of the defects). For GUT scale phase transitions $l \sim 10^{-30}$cm as compared to cosmic distances of the order of Mpc $\sim 10^{24}$cm. If we are willing to loose the information of the precise field configuration over these tiny regions, it seems well justified to fix $\phi$ to the vacuum manifold $\cal N$. Instead of discussing the field equation Eq.~(\ref{phi1}), we require $ \phi/\eta \in \bm{ S}^{(N-1)}$. The remaining field equation, $\Box\phi=0$, then demands that \[ \phi/\eta\equiv \beta ~:~ {\cal M} \rightarrow \bm{ S}^{(N-1)} \] is a harmonic map from spacetime $\cal M$ into {\bf S}$^{(N-1)}$. The topological defects we are interested in are singularities of these maps. When the gradients of $\phi$ become very large, like, e.g., towards the center of a global monopole, the field leaves the vacuum manifold and assumes non vanishing potential energy. If $\beta\in \bm{ S}^{(N-1)}$ is enforced, a singularity develops by topological reasons. In the physics literature harmonic maps are known as $\sigma$--models. The action of a $\sigma$--model is given by \begin{equation} S_\sigma = \int_{\cal M}g^{\mu\nu}\partial_\mu\beta^A\partial_\nu\beta^B\gamma_{AB}(\beta) \sqrt{|g|}d^4x ~, \label{Ssi} \end{equation} where $\gamma_{AB}$ denotes the metric on $\bm{ S}^{N-1}$ and $g_{\mu\nu}$ is the metric of spacetime. We now fix $\beta$ to lay in the vacuum manifold, $\bm{ S}^{N-1}$ by introducing a Lagrange multiplier. We then obtain the following equation of motion for $\beta$: \begin{equation} \Box\b - (\b\cdot\Box\b)\b =0 ~, \label{si2} \end{equation} which shows that the $\sigma$--model is scale free. There are thus two possible evolution equations for the scalar field at low temperature. We call Eq.~(\ref{phi1}) the 'potential model' evolution equation and Eq.~(\ref{si2}) the $\sigma$--model approach. The energy momentum tensor of the scalar field perturbs spacetime geometry and induces perturbations in the dark matter energy density according to Eq.~(\ref{C}) \begin{equation} \ddot{D} + ({\dot{a}\over a})\dot{D} - 4\pi Ga^2\rho_{C}D = 8\pi G \dot{\phi}^2 \label{D..} ~, \end{equation} where $D$ is a gauge invariant variable for the dark matter perturbations \cite{d90}. On subhorizon scales $D \sim \delta\rho/\rho$. In comoving coordinates, the total perturbed energy momentum tensor is given by \[ \delta T_\mu^\nu = \phi,_\mu\cdot\phi^{,\nu}-{1\over 2}\delta_\mu^\nu \phi,_\lambda\cdot\phi^{,\lambda} +\rho_{C}D\delta_\mu^0\delta_0^\nu ~.\] As already mentioned in section~2, the perturbed Einstein equations to this energy momentum tensor yield an algebraic equation for the divergence of the electric part of the Weyl tensor and an evolution equation for the magnetic part of the Weyl tensor ({\sl see Appendix~A}): \begin{equation} \partial^jE_{ij} = -{8\pi\over 3} G\rho_{C}a^2D_i - 8\pi G({1\over 3}\partial_i\delta T_{00} +({\dot{a}\over a})\delta T_{0i} + {1\over 2}\partial^j\tau_{ij}) \label{djE} ~,\mbox{ and} \end{equation} \begin{equation} {1\over a}(aB)_{ij})^{\cdot\cd} + -\triangle B_{ij} = 8\pi G{\cal S}^{(B)}_{ij} ~, \label{B..} \end{equation} \[ \mbox{with }~~~ {\cal S}^{(B)}_{ij}= \epsilon_{lm(i}[T^{S}_{0l},_{j)m} + \dot{\tau}_{j)l},_m]~, ~~~\mbox{ and }~~~~ \tau_{ij}=\phi,_i\phi,_j-{1\over 3}\delta_{ij}(\nabla\phi)^2 ~.\] The source term for the perturbation of Liouville's equation is given by Eq.~(\ref{STS}): \begin{equation} -3n^i\partial^jE^{(S)}_{ij} + n^kn^j\epsilon_{klj} \partial_lB^{(S)}_{ij} \equiv S_{ST}(t,\bm{x},\bm{n})~ . \label{sou} \end{equation} The CMB fluctuations are then determined according to \begin{eqnarray} {\delta T\over T}(t,\bm{x},\bm{n}) = \int_{t_i}^t S_{ST}(t',\bm{x}+(t'-t)\bm{n}, \bm{n})dt' ~ + ~ \Phi_S(t_i,\bm{x}+(t_i-t)\bm{n}, \bm{n}) ~, \nonumber \\ + {1\over 3}\Psi_C(t_i,\bm{x}_i) - {2\over 3}\dot{\Psi}_C/(\dot{a}/a)(t_i,\bm{x}_i) +2\int_i^f\dot{\Psi}_C dt \label{chi2} \end{eqnarray} Eqs.~(\ref{phi1}) and (\ref{D..}) to (\ref{chi2}) form a closed, hyperbolic system of partial differential equations. Actually all except the scalar field equation Eq.~(\ref{phi1}) are linear perturbation equations with source term and can thus be solved, e.g., by the Wronskian method, i.e., by some integrals over the source term. The corresponding solution for $\delta T/T$ is given above in Eq.~(\ref{chi2}), the general solution of the dark matter equation is given below in Eqns.~(\ref{D2}), (\ref{Dgen}) and (\ref{wrons}). Let us briefly describe the general solution for $B_{ij}$: We switch to Fourier space, because there the $\triangle$ is a simple multiplication by $-k^2$ and Eq.~(\ref{B..}) becomes an ordinary differential equation with scalar homogeneous solutions \begin{equation} b^{\pm} = {1\over a}\exp(\pm ikt) ~. \end{equation} The general solution to the inhomogeneous equation is given by \begin{equation} B_{ij}=(b^{+}C^{+}_{ij} +b^{-}C^{-}_{ij}) + B^{(hom)}_{ij} ~, \label{BW}\end{equation} where $B^{hom}$ denotes an arbitrary homogeneous solution and $C^{+}$, $C^{-}$ are given by \begin{eqnarray} C^{+}_{ij} &=& -8\pi G\int{\tilde{\cal S}^{(B)}_{ij}b^- \over W}dt\\ C^{-}_{ij} &=& 8\pi G\int{\tilde{\cal S}^{(B)}_{ij}b^+ \over W}dt~. \end{eqnarray} Here $W$ denotes the Wronskian determinant of the solutions which amounts to \begin{equation} W=b^+\dot{b}^--b^-\dot{b}^+ = {2ik\over a^2} ~. \end{equation} \section{Initial Conditions and Numerical Methods} \subsection{The scalar field:} As already shown in the previous section, the equation of motion of the scalar field is given by \begin{equation} g^{\mu\nu}\nabla_\mu\nabla_\nu\phi + {\partial V\over \partial\phi} = 0 ~, \label{phi}\end{equation} where $g^{\mu\nu}$ is the background (unperturbed metric). With $\beta = \phi/\eta$ and $m=\sqrt{\lambda}\eta$, Eq.~(\ref{phi}) yields for $O(N)$ models in a Friedmann universe \begin{equation} \partial_t^2\beta + 2(\dot{a}/a)\partial_t\beta -\nabla^2\beta = {1\over 2}a^2m^2(\beta^2-1)\beta ~. \label{bb} \end{equation} This equation as it stands can not be treated numerically in the regime which is interesting for large scale structure formation. The two scales in the problem are the horizon scale $t\sim (\dot{a}/a)^{-1}$ and the inverse symmetry breaking scale, the comoving scale $(am)^{-1}$. At recombination, e.g., these scales differ by a factor of about $10^{53}$ and can thus not both be resolved in one computer code. There are two approximations to treat the scalar field numerically. As we shall see, they are complementary and thus the fact that both approximations agree with each other within about 10\% is reassuring. The first possibility is to replace $(am)^{-1}$ by $w$, the smallest scale which can be resolved in a given simulation, typically twice the grid spacing, $w \sim 2\Delta x$. The time dependence of $(am)^{-1}$ which results in a steepening of the potential is mimicked by an additional damping term: $ 2(\dot{a}/a) \rightarrow \alpha \dot{a}/a$, with $\alpha \sim 3$ \cite{PSR}. Numerical tests have shown, that this procedure, which usually is implemented by a modified staggered leap frog scheme \cite{NR}, is not very sensitive on the values of $\alpha$ and $w$ chosen. With this method we have replaced the growing comoving mass $am$ by the largest mass which our code can resolve. For a $(256)^3$ grid which simulates the evolution of the scalar field until today, we obtain $256\Delta x \sim t_0 \sim 4\times10^{17}$sec$/a_0$, so that $w\sim 4\times 10^{15}$sec$/a_0$, i.e., $am\sim \eta a_{rec}\sim 10^{17}GeV$ is replaced by about $w^{-1}= a_010^{-39 }GeV \sim 10^{-35}GeV$, where we set $a_{eq}=1$. We are confident that this modified equation mimics the behavior of the field, since the actual mass of the scalar field is irrelevant as long as it is much larger than the typical kinetic and gradient energies associated with the field which are of the order the inverse horizon scale. Therefore, as soon as the horizon scale is substantially larger than $\Delta x$, the code should mimic the true field evolution on scales larger then $w$. But, to our knowledge, there exists no rigorous mathematical approximation scheme leading to the above treatment of the scalar field which would then also yield the optimal choice for $\alpha$. Alternatively, we can treat the scalar field in the $\sigma$--model approximation given in the previous section. This approach is opposite to the one outlined above in which the scalar field mass is much too small, since the $\sigma$--model corresponds to setting the scalar field mass infinity. The $\sigma$--model equation of motion cannot be treated numerically with a leap frog scheme, since it involves non--linear time derivatives. In this case, a second order accurate integration scheme has been developed by varying the discretized action with respect to the field \cite{PST}. The two different approaches have been extensively tested by us and other workers in the field, and good agreement has been found on scales larger than about 3 -- 4 grid sizes \cite{BCL1,B}. We have compared our potential code with the exact spherically symmetric scaling solution \cite{TS} and with our old spherically symmetric $\sigma$--model code \cite{DHZ}. Outside the unwinding events which extend over approximately 3 grid sizes, the different approaches agree within about 5\%. This is very encouraging, especially since the two treatments are complementary: In the $\sigma$--model, we let the scalar field mass $m$ go to infinity. In the potential approach, we replace $m$ by $\sim 1/\Delta x \sim 200/t_{0} \sim 200a_0/10^{10}y \sim 10^{-35}$GeV. The integration of the scalar field equation is numerically the hardest part of the problem, since it involves the solution of a system of nonlinear partial differential equations. A good test of our numerical calculations, next to checking the scaling behavior of $\rho_S$, is energy momentum conservation of the scalar field, $T^{(S)\mu\nu}_{;\nu}=0$. Energy momentum conservation in the potential model, about 15\%, is slightly worse than in the $\sigma$--model, where it is about 5\% ({\sl see Fig.~2}). Therefore, the final results presented here are all obtained with the $\sigma$--model approach. Our checks lead us to the conclusion, that we can calculate the scalar field energy momentum tensor, which then is the source of dark matter and CMB fluctuations to an accuracy of about 10\%. The problem of choosing the correct initial condition may induce another (systematical) error in our calculations which we hope to remain below 20\%. Other sources of error are negligible. \subsection{Dark matter} Once the scalar field $\beta(\bm{x},t)$ is known, the dark matter perturbations can easily be calculated by either using the Wronskian method (see below) or some standard ordinary differential equation solver. We have performed both methods and they agree very well. For later use, we briefly describe the Wronskian method. We normalize the scale factor by \[ a= {t\over \tau}(1+{1\over 4}t/\tau) ~~~, \mbox{ with} \] \[ \tau=1/\sqrt{(4\pi G/3)\rho_{eq}}= {t_{eq}\over 2(\sqrt{2}-1)}~.\] Here $t_{eq}$ denotes the time of equal matter and radiation density, $ \rho_{rad}(t_{eq})=\rho_{C}(t_{eq})=(1/2)\rho(t_{eq})$. We have normalized $a$ such that $a_{eq}=a(t_{eq})=1$. Transformed to the variable $a$, the dark matter equation Eq.~(\ref{C}) then yields \begin{equation} {d^2 D\over da^2} + {2+3a\over 2a(1+a)}{d D\over d a} -{3\over 2a(1+a)}D = 2\epsilon \dot{\beta}^2({ da\over dt})^2 = (1+a)S/\tau^2 ~, \label{da}\end{equation} \[ S=2\epsilon\dot{\beta}^2 ~ \mbox{ and } \epsilon=4\pi G\eta^2 ~.\] The homogeneous solutions to this linear differential equation are well known \cite{GP}: \begin{eqnarray} D_1 &=& 1+{3\over 2}a ~, \label{D1}\\ D_2 &=& (1+{3\over 2}a)[\ln\left({\sqrt{a+1}+1\over\sqrt{a+1}-1} \right) -3\sqrt{a+1}]~. \label{D2} \end{eqnarray} The general solution to Eq.~(\ref{da}) is given by \begin{equation} D(t)=c_1(t)D_1(t) + c_2(t)D_2(t) \label{Dgen} \end{equation} with \begin{equation} c_1 = -\int(SD_2/W)dt \;\;\mbox{ , }\;\; c_2 = \int(SD_1/W)dt \;\; . \label{wrons} \end{equation} \[ W =D_1\dot{D_2} - \dot{D_1}D_2 = {\dot{a}(2a-1)\over a\sqrt{a+1}} = {2a-1\over a\tau} \] is the Wronskian determinant of the homogeneous solutions. The integrals Eq.~(\ref{wrons}) have to be performed numerically with $S=\epsilon\dot{\beta}^2$. When discussing the initial conditions for $D$ in subsection~4.4, we shall present an analytic approximation for the source term $S$. \subsection{The CMB anisotropies} The CMB anisotropies are given by \[ {\delta T\over T} = \triangle^{-1}\chi \] up to monopole and dipole contributions which we disregard. Here $\chi$ is a solution Eq.~(\ref{chi}) of Eq.~(\ref{Lchi}). The source term $S_T$ is determined via Eq.~(\ref{dE}) and Eq.~(\ref{Bij}). However, using this straightforward approach results in a big waste of computer memory (which we cannot afford): We would be satisfied to calculate $\delta T/T$ for about 30 observers in each simulations, which means we need $\triangle^{-1}\chi$ only at 30 positions {\bf x}. But since we have to perform an inverse Laplacian which is done by fast Fourier transforms, we have to calculate $\chi$ on the whole grid, which consists of $192^3 \sim 8\cdot 10^6$ positions. In Addition, to calculate the spherical harmonic amplitudes of $\delta T/T$ up to about $l\sim 40$ (angular resolution of about $4^o$), we need typically $5\times10^4$ directions $\bm{n}$. The $\chi$ variable alone (in double precision) would thus require 16 G-bytes of memory, an amount which is not available on most present day computers. The way out is to take the inverse Laplacian already in the equation of motion Eq.~(\ref{Lchi}). This results in \begin{equation} (\partial_t +\gamma^i\partial_i){\delta T\over T} = -3\gamma^i\triangle^{-1}(\partial^jE_{ij}) - \gamma^k\gamma^j\epsilon_{kli}\triangle^{-1}(\partial_lB_{ij}) \equiv \triangle^{-1}S_T(t,\bm{x},\bm{\gamma})~ . \label{LT} \end{equation} Here the inverse Laplacian has to be performed for a vector field and a symmetric traceless tensor field, a total of 8 scalar variables which only depend on {\bf x} and not on $\bm{\gamma}$. For a $192^3$ grid this requires nearly 1 Gbyte of memory, no problem for presently available machines. Eq.~(\ref{LT}) has the general solution ({\sl see} Eq.~(\ref{genu})) \begin{eqnarray} {\delta T\over T}(t_0,\bm{x}_0,\bm{\gamma}) &=& +\Phi^S(t_i,\bm{x}_i) - 3\int_i^f\triangle^{-1}S_{TS}(t,\bm{x}_0-(t_f-t)\bm{n},\bm{n}) dt \nonumber \\ && + {1\over 3}\Psi_C(t_i,\bm{x}_i) - {2\over 3}\dot{\Psi}_C/(\dot{a}/a)(t_i,\bm{x}_i) +2\int_i^f\dot{\Psi}_C dt ~,\label{dT} \end{eqnarray} with $S_{TS}$ given in Eq.~(\ref{STS}). The first term of Eq.~(\ref{dT}) determines the initial condition of the CMB anisotropies due to the source term. In the numerical simulation we just set it $ -3n^l\triangle^{-1}(\partial^jE^{(S)}_{lj})(t_i,\bm{x}_i)t_i$. This assumes that the source term is approximately constant until $t_i$ and that magnetic contributions can be neglected. The resulting amplitude is not very sensitive to this assumption, but changing it can somewhat change in the spectral index. We have solved Eq.~(\ref{dT}) numerically by just summing up the contributions from each time step for 27 observer positions $\bm{ x}_0$. The value of the source term at position $\bm{x}_0+(t-t_0)$ is determined by linear interpolation. The quantity $\partial^iE^S_{ij}$ is determined by Eq.~(\ref{dE}) and its inverse Laplacian is calculated by fast Fourier transforms. To obtain $\triangle^{-1}B^S_{ij}$ from equation Eq.~(\ref{Bij}), we directly calculate $\triangle^{-1}{\cal S}_B$ in $k$--space, then solve the ordinary, linear differential equation for $\triangle^{-1}B_{ij}$ in $k$--space by the Wronskian method. Since, as we shall argue later, all components $T^{(S)}_{\mu\nu}$ in average scale like $A/\sqrt{t}$ on super horizon scales, $S^(B) \sim At^{-3/2}$ and therefore $C^{\pm} \propto t^{3/2}$ on super horizon scales. Therefore, we can neglect the contribution to $C^\pm$ from the lower boundary in the integral. Furthermore, since the homogeneous solution $B_{ij}^{(hom)}$ is decaying, we drop it entirely. This procedure corresponds to setting $B(t_{i})=0$ and calculating $B(t)$ according to Eq~(\ref{BW}). \subsection{Initial conditions} Initially, the field $\phi$ itself and/or the velocities $\dot{\phi}$ are laid down randomly on the grid points. The initial time, $t_{in}$ is chosen to be the grid size, $t_{in} = \Delta x$, so that the field at different grid points should not be correlated. The configuration is then evolved in time with one of the approximation schemes discussed above. Because our initial conditions for dark matter and photons very sensitively depend on the scaling behavior of the dark matter, we can only start the dark matter or photon simulations when scaling is fully reached, $t_{in}=8\Delta x$. Starting our simulations, e.g., at $t=4\Delta x$ changes the results by about a factor of 2. Further doubling of the initial time, changes our results by less than 20\%, we thus believe that at $t=8\Delta x$ scaling is sufficiently accurately. Unfortunately, this late initial time reduces our dynamical range to about $192/8=24$ for a $192^3$ grid, which is seen clearly in our results for the CMB anisotropies discussed below. It is very important to choose the correct initial conditions for the dark matter and, especially, photon perturbations. Changing them can change the CMB fluctuation amplitudes by nearly a factor two. Since these fluctuations are used to normalize the model, i.e. to determine $\epsilon$, this reflects in corresponding changes in $\epsilon$. We want to do better than a factor two by choosing physically plausible initial conditions. The cleanest way would by to simulate the evolution of perturbations through the phase transition, assuming that before the phase transition, the universe was an unperturbed Friedmann universe with $\phi\equiv 0$. On the other hand, since we want to calculate the perturbation spectrum on scales of up to 1000Mpc with a $(256)^3$ grid, we cannot start our dark matter and CMB simulation earlier than at a time when the horizon distance is approximately $8\Delta x \sim 30$Mpc. At the beginning of the scalar field simulation our grid scale $\Delta x\sim 4$Mpc is of the order of the horizon scale. We therefore have to decide on the amplitudes of super horizon perturbations. One possibility is setting all geometrical perturbations initially to zero. The requirements \[ \partial^jE_{ij}(t_{i})=0 ~~\mbox{ and }~~~ S^B_{ij}(t_{i})=0\] then yield initial conditions for the dark matter fluctuations $D$ and the photon variable $\chi$. But these, let us call them 'strict isocurvature' initial conditions are not natural since they do not propagate in time: Even if we start with $E$ and $B$ vanishing on super horizon scales, after some time residual fluctuations have leaked into these scales and one obtains the white noise fluctuations spectrum on super horizon scales shown in {\sl Fig.~3}. This does not violate causality, since white noise is uncorrelated and just results from the residuals of correlated fluctuations on smaller scales. The correct initial values for $D$ and $\dot{D}$ would of course be those obtained by solving the equation of motion Eq.~(\ref{C}) from the symmetry breaking time until the start of the simulation. We found a method to incorporate this at least approximately: The spectrum of the dark matter source term $8\pi G| \widetilde{\dot{\phi}^2}|^2 $ can be approximated by \begin{equation} 8\pi G |\widetilde{\dot{\phi}^2}| =2\epsilon \widetilde{\dot{\beta}^2} = \epsilon\sqrt{1\over V}\int d^3x\dot{\beta}^2(x)e^{ikx} \approx {\epsilon A\over \sqrt{t}(1+a_1kt +a_2(kt)^2)} ~, \label{fit}\end{equation} with \[ A= 3.3 ~,~~~ a_1=-0.7/(2\pi)~,~~~ a_2 = 0.7/(2\pi)^2 ~. \] This numbers have been obtained by a $\chi^2$--minimization scheme. The approximation is not very good. It yields a $\chi^2 \approx 2000$ for about 1000 data points. Its comparison with the real data in {\sl Figs.~4 and 5} shows that Eq.~(\ref{fit}) approximates the source term to about 10\% on superhorizon scales, but does not follow the wiggles present in the data on smaller scales. Since we shall not use the fit on subhorizon scales, this is not important for our simulations. However, in general $\widetilde{\dot{\phi}^2}$ is complex and setting it equal to its absolute value, we neglect the evolution of phases. Again, by causality, this will not severely affect scales larger than the horizon, since on these scales the phases are (approximately) frozen. But on subhorizon scales our fis is nor very useful due to the incoherent evolution of phases. Assuming this form of the source term, we can solve Eq.~(\ref{C}) analytically on super horizon scales, where we approximate the source term by \begin{equation} 2\epsilon \widetilde{\dot{\beta}^2} = {\epsilon A\over \sqrt{t}}, ~~~~~\mbox{ on super horizon scales.} \label{sqrtt}\end{equation} The homogeneous solutions of Eq.~(\ref{C}) are given by Eq.~(\ref{D1},\ref{D2}) The general inhomogeneous solution, $D=c_1D_1 +c_2D_2$, even with the simple source term Eq.~(\ref{sqrtt}), becomes rather complicated. But in the radiation and matter dominated regimes we find the simple approximations \begin{eqnarray} D &=& (4/7)t^2S ~~;~~ \dot{D}=(6/7)t^2S ~~~ \mbox{ radiation dominated} \label{Dinr} \\ D &=& -(4/9)t^2S ~~;~~ \dot{D}=-(2/3)t^2S ~~~\mbox{ matter dominated.} \label{Dinm}\end{eqnarray} From $D$ we can calculate $\Psi_C$, leading to the dark matter contribution to the CMB anisotropies. As mentioned above, the initial contribution of the scalar field is approximated by \[ \Phi^S(t_i,\bm{x}_i) \sim -3t_in^i\triangle^{-1}(\partial^jE^{(S)}_{ij}) ~.\] The result is very sensitive to initial conditions: If we do not separate the dark matter and choose some arbitrary, non-adiabatic initial condition, the resulting $C_\ell$'s increase by nearly a factor of $10$ and the dark matter induces $80\%$ of the total fluctuation. However, choosing the adiabatic initial condition discussed in Section~3, leading to Eq.~(\ref{dT}), dark matter only contributes about 20\% to the $C_\ell$'s and the main contribution is due to the defects. The dark matter contribution to the CMB anisotropies is not scale invariant, but white noise. It has spectral index $n=0$. This result was found numerically ({\sl see Fig.~6}) but, as we argue in Appendix~B, it can also be understood analytically. our value of $\epsilon$ obtained with these physical isocurvature and on super horizon scales adiabatic initial conditions is in reasonable agreement with the values obtained in \cite{PST} and \cite{BR}. Let us also present a heuristic derivation of the numerical finding Eq.~(\ref{sqrtt}) on superhorizon scales: We know that the average value $\langle\dot{\beta}^2\rangle \propto 1/t^2$, the usual scaling behavior. The Fourier transform of $\dot{\beta}^2$ determines the fluctuations on this 'background' on a given comoving scale $\lambda=2\pi/k$. As long as this scale is super horizon, $\lambda>t$, a patch of size $\lambda^3$ consists of $N=(\lambda/t)^3$ independent horizon size volumes. The fluctuations on this scale should thus be proportional to \[\tilde{\dot{\beta}^2}\propto \langle\dot{\beta}^2\rangle /\sqrt{N} \propto 1/\sqrt{t} ~, \] which is just the behavior which we have found numerically on super--horizon scales. As soon as a given scale becomes sub--horizon, $\lambda\ll t$, $\tilde{\dot{\beta}^2}$ starts decaying from this large scale value like $1/t^2$. \section{Results} \subsection{CMB anisotropies} To analyze the CMB anisotropies, we expand $\delta T/T$ in spherical harmonics \begin{equation} {\delta T\over T}(t_0,\mbox{\boldmath{$x,\gamma$}}) = \sum_{lm}a_{lm}(\mbox{\boldmath{$x$}})Y_{lm}(\mbox{\boldmath{$\gamma$}}) ~. \end{equation} As usual, we assume that the average over $N_x$ different observer positions coincides with the ensemble average and define \begin{equation} C_\ell = {1\over (2\ell+1)N_x} \sum_{m,x} |a_{\ell m}(\mbox{\boldmath{$x$}})|^2 ~, ~~ \ell\ge 2 \label{cl} ~. \end{equation} Gaussian fluctuations are characterized by the two point correlation function. Since the angular two point correlation function is given by \begin{equation} \langle {\delta T\over T}(\bm{ n}){\delta T\over T}(\bm{ n}') \rangle_{(\bm{ n\cdot n}'=\cos\theta)} ={1\over 4\pi} \sum_\ell(2\ell+1) C_\ell P_\ell(\cos\theta) ,\end{equation} Gaussian distributed CMB fluctuations are fully determined by the $C_\ell$'s. However, as can be seen from Fig.~7, in our case the distribution of the CMB fluctuations is not quite Gaussian. It is slightly negatively skewed. We find an average skewness of $-0.5$ and a kurtosis of $0.7$. For a simulation on a $192^3$ grid with 27 different observer positions for each simulation. The harmonic amplitudes with are shown in Fig.~8. The low order multipoles depend strongly on the random initial conditions (cosmic variance), like in the spherically symmetric simulation \cite{DHZ}. It is well known, that cold dark matter fluctuations with a power spectrum of spectral index $n$ gravitationally induce CMB anisotropies with a spectrum given by \cite{Ef} \begin{equation} C_\ell = C_2{\Gamma(l+(n-1)/2)\Gamma((9-n)/2)\over \Gamma(l+(5-n)/2)\Gamma((n+3)/2)} \label{spec} ~.\end{equation} We have performed a least square fit of our numerical results for $\log(C_\ell)$ and the $\log(C_\ell)$ obtained from Eq.~(\ref{spec}).\footnote{In the case of topological defect induced fluctuations, the $C_\ell$ spectrum does not have exactly this form, since CBM fluctuations are not only induced by the dark matter but mainly by the scalar field perturbations and the assumptions made for the derivation of this formula are not valid.} If we take into account all the $C_\ell$'s reliably calculated in our simulations, which limits us approximately to $\ell\le 22$ we find a very nice scale invariant spectrum, \begin{equation} n= 0.9\pm 0.2 \end{equation} with quadrupole amplitude \begin{equation} Q =\sqrt{(5/4\pi)C_2}T_{CMB} = 2.8\pm 0.7 K \cdot \epsilon ~. \label{Q}\end{equation} The 1, 2 and 3 sigma contour plot is shown in Fig.~9. The minimal $\chi^2$ is 0.56. It is very interesting, that the dark matter contribution to the CMB anisotropies does not yield a scale invariant spectrum, but white noise. This can be understood analytically: The $\dot{\Psi}_C$ contributions to $\delta T/T$ in Eq.~(\ref{genu}) are not very important and \[ (\delta T/T)_C(t_0,\bm{k}) \sim {1\over 3}\Psi_C(t_i,\bm{k})\exp(i\bm{k\cdot nt_0}) = {\epsilon A\over 6\sqrt{t_i}}{\exp(i\bm{k\cdot nt_0})\over k^2} \] on super horizon scales. For the second equal sign we used $k^2\Psi_C=4\pi GD_C\sim {\epsilon A\over 2\sqrt{t_i}}$. By standard arguments ({\sl see, e.g.} \cite{Ef}) one then finds \begin{equation} C_\ell^{(C)} = {\epsilon^2A^2 \over 18\pi t_i}\int{dk\over k^2}j_\ell^2(kt_0) \propto {\Gamma(\ell-0.5)\over \Gamma(\ell+2.5)} ~, \end{equation} corresponding to Eq.~(\ref{spec}) with $n=0$. This is also what we find numerically ({\sl see Fig.~6}). The dark matter contribution caused the spectral index $n$ of the total CMB anisotropies to drop slightly below $n=1$. To reproduce the COBE amplitude $ Q_{COBE}= (20 \pm 5)\mu K$ \cite{Go}, we have to normalize the spectrum by choosing the phase transition scale $\eta$ according to \begin{equation} \epsilon = 4\pi G\eta^2 = (0.8 \pm 0.4)10^{-5} \label{ep}~. \end{equation} This value is somewhat smaller, but still comparable with the results obtained in \cite{BR,PST}. Another method to determine $\epsilon$ is the following: The total temperature fluctuation amplitude on a given angular scale $\theta_c$ is given by \begin{equation} \sigma^2_T(\theta_c) = {1\over 4\pi} \sum_\ell C_\ell(2\ell+1)\exp(-\ell^2\theta_c^2/2) ~. \end{equation} In Fig.~10 we show $\sigma_T$ as a function of $\theta_c$. In a recent analysis of the COBE data\cite{WB} $\sigma_T^{(COBE)}(7^o) \sim 44\mu K$ and $\sigma_T^{(COBE)}(10^o) \sim 40\mu K$ for a spectral index $n\sim 1$, which leads again to the result given in Eq.~(\ref{ep}). \subsection{Dark matter fluctuations} Using fast Fourier transforms we calculate the spectrum $P(k) = |\delta(k)|^2$ of the dark matter density fluctuations is shown in Fig.~11. The fit shown as solid line in Fig.~11 is given by \begin{equation} P(k)h^3/(2\pi)^3= {Ck\over (1 +\alpha k + (\beta k)^{1.5} +(\gamma k)^2)^2} ~, \label{Pk} \end{equation} with $h=0.5$ and \begin{eqnarray} C &=& 215h^{-1}Mpc^4 \label{CC}\\ \alpha &=& 10 h^{-2}Mpc ~~(=0.5\tau) \\ \beta &=& 1.25h^{-2} Mpc ~~ \sim \tau/(4\pi) \\ \gamma &=& 2.3 h^{-2} Mpc ~~ \sim \tau/(2\pi) ~, \end{eqnarray} where we have used $\tau=19.36h^{-2}$Mpc which is approximately the comoving time at equal matter and radiation. The parameter $C$, which is most important to determine the bias factor can also be obtained by the following rough analytical argument: On super horizon scales, $|D|^2 \sim (0.5\epsilon A)^2t^3$ according to Eq.~(\ref{Dinm}). As soon as the perturbation enters the horizon at $t=2\pi/k$, the source term disappears and $D$ starts growing like $t^2$, leading to \begin{equation} P(k,t_0) \sim {(0.5\epsilon A)^2\over 2\pi}kt_0^4 = {(2\pi)^3\over h^3}C_{an}k ~. \end{equation} Inserting the numbers $\epsilon=0.8\times 10^{-5}$, $A=3.3$, $t_0^2 =4a_0\tau^2$, $ a_0\sim 2.5h^2\times 10^4$, we obtain $C_{an}\sim 190$ in excellent agreement with Eq.~(\ref{CC}). Figure~11 can be compared directly with the IRAS observation \cite{IRAS} and it is compatible with a bias factor of order 1. A more detailed calculation with Gaussian or square hat window function yields for $\epsilon=0.8\times 10^{-5}$ \begin{eqnarray} \sigma_{sim}(10MPc)= 1/b_{10} \sim 0.5 - 1 & \sigma_{QDOT}(10MPc)= 1& \mbox{for $h=0.5$}, \end{eqnarray} yielding $b_{10}\sim 1 - 2 $ for the value of $\epsilon$ found by comparison with COBE Eq.~(\ref{ep}) A value even somewhat closer to 1 is found for $b_{20}$. Observations and simulations of nonlinear clustering of dark matter and baryons \cite{CO} suggest a bias factor $b_{10}\sim 1 - 2$ which is compatible with our results. It is remarkable, that unlike in the simulations by Pen et al. \cite{PST}, our bias factor is approximately constant and physically acceptable. (To determine our power spectrum, we have not taken into account any smoothing which might change the results by at most 15\%.) In Fig.~12 we have shown the dark matter pixel distribution from a $100^3$ simulation. It is interesting, that the skewness of the dark matter distribution is positive, where the $\delta T/T$ skewness is negative. \section{Conclusions} Our simulations show that global texture lead to a scale invariant spectrum of microwave background fluctuations on large scales like inflationary models of structure formation. This is one of the main results of this investigation. It is however interesting that the dark matter contribution to the CMB anisotropies is not scale invariant, but white noise. Therefore it is important that the initial condition for dark matter and radiation are adiabatic in which case the dark matter contribution to the $C_\ell$'s is small and the flat spectrum caused by the defects is maintained. Our second main result is the dark matter fluctuation spectrum. The spectrum is very close to scale invariant and the bias factor needed for $\epsilon$ from the CMB anisotropies is around $b\sim$ 1--2. This value is certainly acceptable and smaller than the bias factor obtained in previous investigations \cite{CST}. The deviation from Gaussian statistics seems to us not very significant ({\sl see Figs.~7 and 12}) and it is thus important to develop other means to distinguish topological defects from inflationary scenarios. A clean and promising candidate for this distinction are the Doppler peaks which are calculated for the texture scenario in \cite{DGS}. From our investigations we thus conclude that concerning the large scale CMB anisotropies and the linear dark matter perturbation spectrum the texture scenario and probably also other models with global defects are compatible with present observations. \vspace{2cm} \noindent {\large\bf Acknowledgment} We thank the staff at CSCS, for valuable support. Especially we want to mention Andrea Bernasconi, Djiordic Maric and Urs Meier. We acknowledge discussions with Joachim Laukenmann who carefully read the manuscript and checked the more involved algebraic derivations with Maple. We profited from many helpful and encouraging discussions of this work especially we want to mention Marc Hindmarsh, Philipp Jetzer, Yipeng Jing, Mairi Sakellariadou, Norbert Straumann, Neil Turok, Simon White and others. We also thank Neil Turok for providing us their $\sigma$--model code.
\section{Introduction} Let F to be $R,C,H$.The inner product on $F^{n+1}$ is given as; $$<(z_1,\cdots,z_{n+1}),(w_1,\cdots,w_{n+1})>= \sum_1^n{z_i\bar{w_i}}-z_{n+1}\overline{w_{n+1}}$$ Let $GL(n+1,F)$ act on $F^{n+1}$ from the right and let $P^n_F$ be the space of left $F$-lines. Then $G=O_F(n,1)$ is a subgroup of $GL(n+1,F)$ preserving this inner product. The induced action of $O_F(n,1)$ on $P^n_F$ gives three invariant subsets $D_+,D _0,D_-$ and the action is positive definite on $D_+$.A component of $D_+$ is called hyperbolic space.An isotropy group of $O_F(n,1)$ is a maximal compact subgroup $K=O_F(1) \times O_F(n)$.If we numerate them; $$ O_R(n,1)=O(n,1),K=O(1)\times O(n)$$ $$ O_C(n,1)=U(n,1),K=U(1)\times U(n)$$ $$ O_H(n,1)=Sp(n,1),K=Sp(1) \times Sp(n)$$ The hyperbolic Cayley plane is similarly realized and the isometry group is the real form of $F_4$ of real rank one.The maximal compact subgroup is $Spin(9)$. We can compactify the hyperbolic space by adding $D_0$.If $dim_RF=k$ then $D_0$ is a $(kn-1)$ dimensional sphere denoted by $\partial{H^n_F}$. Take the Iwasawa decomposition $G=KAN$.Then $A$ and $N$ fix $\xi$ in the sphere at infinity and $N$ is a nilpotent goup obtained as an extion $$ 0 \rightarrow ImF \rightarrow N \rightarrow F^{n-1} \rightarrow 0.$$ Horospheres based at $\xi$ are the orbits of $N$. The stabilizer in $O_F(n,1)$ of $\xi$ is a parabolic subgroup with Langlands decomposition $MAN$.For $F=R,C,H$ $M$ is a subgroup of $K$ of the form $O_F(1)\times O_F(n-1)$.For Cayley plane $M$ is $Spin(7)$. $K$ actually acts transitively on $\partial{H^n_F}$ giving $K/M=\partial{H^n_F}$ .The centralizer of $M$ in $K$ acts freely on $K/M$ with quotient $P^{n-1}_F$. This gives Hopf fibration $\partial{H^n_F}\rightarrow P^{n-1}_F$. Since $K$ acts transitively on $\partial{H^n_F}$ ,killing form on its Lie algebra determines a Riemannian metric on $\partial{H^n_F}$.Define the orthogonal complements to the vertical tangent spaces of the Hopf fibration $\partial{H^n_F} \rightarrow P^{n-1}_F$.Then there is $K$ invariant Riemannian metric on the horizontal bundle and if we define the distance of two points as the minimum of the lengthes of the pathes joining these two points and staying horizontal bundle it is Carnot-Caratheodary metric.Mitchell[Mi] calculated that the housdorff dimension of $\partial{H^n_F}$ in this metric is $k(n+1)-2$. \section{Boundary of Rank one symmetric Spaces} The boundary of rank one symmetric space $H^m_F$ is a one point compactification of nilpotent group $N$ in Iwasawa decomposition,denoted by $N \cup \infty$.We will introduce nice coordinates and left invariant distance to introduce cross ratio of four points.We will define everything in Cayley numbers because that is the most general case among four. Cayley number is a pair of quaternions $(q_1,q_2)$.The multiplication is defined as: $$(q_1,q_2)(p_1,p_2)=(q_1p_1-\bar{p_2}q_2,q_2p_1+p_2\bar{q_1})$$ Also we define $\overline{(q_1,q_2)}=(\bar{q_1},-q_2)$. Then it has the following properties. $1)q\bar{q}=|q|^2$ $2)|qp|=|q||p|$ $3)q^{-1}=\bar{q}/|q|^2$ $4)\overline{pq}=\bar{q}\bar{p}$ Even though Cayley numbers are not commutative nor associative,using Artin's theorem saying that the subalgebra generated by two elements is associative,we have following extra properties. 1)$(xy)y^{-1}=x$ 2)$(xy)^{-1}=y^{-1}x^{-1}$ Here is some usuful fact. $Aut(O)=G_2$ and it is compact group.It fixes real number and acts transitively on unit pure imaginaries.The stabilizer of $i$ is a copy of $SU(3)$ and it acts transitively on the unit pure imaginaries orthogona l to $i$.The stabilizer of $i,j$ which fixes $k$ since $k=ij$,acts transitively on the unit pure imaginaries orthogonal to $i,j,k$ and it is a copy of $SU(2)$. For more extensive informations,see [Fr]. Now an element of $N$ is denoted as $[(t,q),k]$ where 1)$F=R(Real),(t,q)=0,k \in R^{m-1}$. 2)$F=C(Complex),t$ is real,$q=0,k \in C^{m-1}$. 3)$F=H(Quaternion),t \in R^3,q=0,k \in H^{m-1}$. 4)$F=O(Cayley),t \in R^3,q$ is a quaternion,k is a Cayley number. The multiplication is defined as; $$[(t,q),k][(t',q'),k']=[(t+t',q+q')+2Im<k,k'>,k+k']$$ We define the gauge of $[(t,q),k]$ as; $$A([(t,q),k])=(|k|^2+t,q)$$ Then define $|[(t,q),k]|=(|k|^4+|t|^2+|q|^2)^{1/4}$ and $d(g,g')=|g'^{-1}g|$. This is a left invariant distance and the housdorff dimesion of the boundary with respect to this metric is $dim_R(F) \times (m+1)-2$ wich agrees with Mitchell's calculation.Even though this metric is not riemannian if we take the inner metric of this it becomes Carnot-Caratheodory metric.For references see[Gr][Pa][Mi]. The action of $Iso(H^m_F)$ extends continously to $\partial(H^m_F)$.Let $Sim(N)$ denote the subgroup of $Iso(H^m_F)$ wich fixes $\infty$.Then it is isomorphic to $ N \rtimes (O_F(m-1)\cdot O_F(1) \times R)$ and $N \rtimes O_F(m-1)\cdot O_F(1)$ acts as isometries with respect to the metric given above.Note that $N$ part comes from left action of the group itself and $O_F(m-1)\cdot O_F(1) \times R$ part comes from hyperbolic isometries of the form in $O_F(m,1)$ \[ \left[ \begin{array}{ccc} M & 0 & 0 \\ 0 & \nu\cosh(s) & \nu\sinh(s) \\ 0 & \nu\sinh(s) & \nu\cosh(s) \end{array} \right] \] where $O_F(m-1)\cdot O_F(1)=O(m-1),U(m-1),Sp(m-1)\cdot Sp(1),O_O(1)\cdot O_O(1)$ depending on $F=R,C,H,O$ respectively.These actions will be described in next section. We define the cross ratio of four points as; $$[g_1,g_2,g_3,g_4]=\frac{|A(g_3^{-1}g_1)||A(g_4^{-1}g_2)|} {|A(g_4^{-1}g_1)||A(g_3^{-1}g_2)|}$$ Let's diverge to the unit ball model for a while so that we can get some connetion between the metric of hyperbolic space and the metric defined above. In Real,complex,Quaternion case for $x,y \in H^m_F$ $$cosh(d(x,y))=\frac{|1-<x,y>|}{(1-<x,x>)^{1/2}(1-<y,y>)^{1/2}}$$ For Cayley hyperbolic case if we define $$R<v,w>=Re(v_1\bar{v_2})(w_2\bar{w_1})-Re(\bar{v_2}w_2)(\bar{w_1}v_1)$$ Then for $x,y \in H^2_O$ the distance is $$cosh(d(x,y))=\frac{(|1-<x,y>|^2+2R<x,y>)^{1/2}} {(1-<x,x>)^{1/2}(1-<y,y>)^{1/2}}$$ For details see[Mo]. Let $<<x,y>>=1-<x,y>$ for $F=R,C,H$ and $<<x,y>>={(|1-<x,y>|^2+2R<x,y>)^{1/2}}$ for Cayley case.Then the cross ratio of the four points in the boudary of hyperbolic space is defined as $$[x,y,z,w]=\frac{<<z,x>><<w,y>>}{<<w,x>><<z,y>>}$$ Note that this is a limit of $$\frac{\cosh(d(z_i,x_i))\cosh(d(w_i,y_i))} {\cosh(d(w_i,x_i))\cosh(d(z_i,y_i))}$$ where $x_i,y_i,z_i,w_i$ tend to $x,y,z,w$.So it is invariant under isometries of hyperbolic space. To make connection between these two definitions of cross ratio we introduce generalize projection from $N \cup \infty$ to the boundary of unit ball. $$w_1=2(1+|k|^2-t,-q)^{-1}k$$ $$w_2=(1+|k|^2-t,-q)^{-1}(1-|k|^2+t,q)$$ Note that $(0,0)$ corresponds to $(0,1)$,$\infty$ to $(0,-1)$. Let $g_i=[(t_i,q_i),(c_i,d_i)]$ for $i=1,2$. By the general projection these two points correspond to $$x_i=1/((1+|k_i|^2)^2+|t_i|^2+|q_i|^2) [2(c_i+|k_i|^2c_i+t_ic_i-\bar{d_i}q_i,q_ic_i+d_i+$$ $$d_i|k_i|^2-d_it_i),(1-|k_i|^4+2t_i-|t_i|^2-|q_i|^2,2q_i)]$$ Since (0,0) and $\infty$ correspond to $(0,1)$ and $(0,-1)$ $$<<x_i,(0,-1)>>=2/((1+|k_i|^2)^2+|t_i|^2+|q_i|^2)$$ $$<<x_i,(0,1)>>=2\frac{\sqrt{(|k_i|^4+|k_i|^2+|t_i|^2+|q_i|^2)^2 +|t_i|^2+|q_i|^2}}{(1+|k_i|^2)^2+|t_i|^2+|q_i|^2}$$ Since $((|k_i|^4+|k_i|^2+|t_i|^2+|q_i|^2)^2+|t_i|^2+|q_i|^2=(|k_i|^4+|t_i|^2+|q_ i|^2)((1+|k_i|^2)^2+|t_i|^2+|q_i|^2)$ we get $$[\infty,0,x_1,x_2]=\frac{(k_2|^4+|t_2|^2+|q_2|^2)^{1/2}} {(k_1|^4+|t_1|^2+|q_1|^2)^{1/2}}$$ $$=|A(g_2^{-1})|/|A(g_1^{-1})|$$ so we are done. For the case $[0,g_1,\infty,g_2]$,just note that $[0,g_1,\infty,g_2]=[g_2^{-1},g_2^{-1}g_1,\infty,0]$.By using above fact we are done again. In fact for $F=R,C,H$ it is easy to show that two definitions of cross ratio agree by brutal calculation but in Cayley Hyperbolic case calculation is quite hard.But to prove our main theorem we will just need what we proved. \section{Action of Isometries on $\partial{H^m_F}$} In this section we want to incode the action of hyperbolic isometry on the boundary.A hyperbolic isometry in $O_F(m,1)$ has the form of ; \[ \left[ \begin{array}{ccc} M & 0 & 0 \\ 0 & \nu\cosh(s) & \nu\sinh(s) \\ 0 & \nu\sinh(s) & \nu\cosh(s) \end{array} \right] \] where $\nu=1,M \in O_F(m-1)$ for $F=R,C$ and $|\nu|=1,M \in Sp(m-1)$ for $F=H$ and $ |M|=|\nu|=1$ for Cayley hyperbolic case. Then this element send $[(t,q),k]$ to; $$(t',q')=e^{-2s}{\nu}^{-1}(t,q)\nu$$ \begin{eqnarray*} k'=e^{-s}\{[{\nu}^{-1}(e^{2s}+|k|^2-t,-q)^{-1}{\nu}]^{-1}\}& & \\ \{[{\nu}^{-1}((e^{2s}+|k|^2-t,-q)^{-1}(1+|k|^2-t,-q))][((1+|k|^2-t,-q)^{-1}k)A] \}& & \end{eqnarray*} First note that the matrix acts on the unit ball as; $$w_1'=(w_2{\nu}\sinh(s)+{\nu}\cosh(s))^{-1}(w_1A)$$ $$w_2'=(w_2{\nu}\sinh(s)+{\nu}\cosh(s))^{-1}(w_2\nu\cosh(s)+\nu\sinh(s))$$ using coordinate chage between unit ball model and $F^{m+1}$ sending $w$ to $(w,1)$. To check above claim we use Artin's theorem crucially. By using generalized projection and above equations we can show \begin{eqnarray*} w_2'=\{((1+|k|^2-t,-q)^{-1}(1-|k|^2+t,q))\nu\sinh(s)+\nu\cosh(s)\}^{-1}& & \\ \{[(1+|k|^2-t,-q)^{-1}(1-|k|^2+t,q)]\nu\cosh(s)+\nu\sinh(s)\}& & \end{eqnarray*} $$=[{\nu}^{-1}((e^{2s}+|k|^2-t,-q)^{-1}(1+|k|^2-t,-q))] [((1+|k|^2-t,-q)^{-1}(e^{2s}-|k|^2+t,q))\nu]$$ equal to $$[({\nu}^{-1}(e^{2s}+|k|^2-(t,q))^{-1})\nu] [{\nu}^{-1}(e^{2s}-|k|^2+(t,q)\nu]$$ To prove this ,set $r_1=e^{2s}+|k|^2,r_2=e^{2s}-|k|^2,r_3=1+|k|^2,(t,q)=Q$ and using identity $-QQ=|Q|^2$ we can show $$[{\nu}^{-1}((r_1+Q)^{-1}(r_3-Q))][((r_3+Q)(r_2+Q))\nu]= [({\nu}^{-1}(r_1-Q)^{-1})\nu][{\nu}^{-1}(r_2+Q)\nu].$$ For $w'_1$,it is easy to check. \section{Marked length spectrum determines representation} First we will make very simple but important observation to prove the theorem. \begin{lemma}Let $a,b$ be two hyperbolic isometries such that $x_1$ is a repelling fixed point of $a$ ,$x_3$ is a attracting fixed point of $a$ and $x_4$ is a attracting fixed point of $b$,$x_2$ is a repelling fixed point of $b$.Then $$\lim_{n\rightarrow\infty}e^{l(a^n)+l(b^n)-l(a^nb^n)}=|[x_1,x_2,x_3,x_4]|.$$ \end{lemma} \begin{pf}Choose $x_1^n$ on the axis of $a$,$z_1^n$ on the axis of $b^na^n$,$x_2 ^n$ on the axis of $b$,$w_2^n$ on the axis of $a^nb^n$ so that $d(x_1^n,z_1^n)$ and $d(x_2^n,w_2^n)$ go to zero.If we put $x_3^n=a^n(x_1^n),z_3^n=a^n(z_1^n),x_4 ^n=b^n(x_2^n),w_4^n=b^n(w_2^n)$ and $d_{13}^n=d(x_1^n,x_3^n),d_{24}^n=d(x_2^n,x_ 4^n),d_{23}^n=d(w_2^n,z_3^n), d_{14}^n=d(z_1^n,w_4^n)$ then $$\lim \sqrt{\frac{<<x_1^n,x_3^n>><<x_3^n,x_1^n>> <<x_2^n,x_4^n>><<x_4^n,x_2^n>>} {<<x_1^n,x_4^n>><<x_4^n,x_1^n>><<x_2^n,x_3^n>><<x_3^n,x_2^n>>}}$$ $$=\lim \frac{(e^{d^n_{13}}+e^{-d^n_{13}})(e^{d^n_{24}}+e^{-d^n_{24}})}{(e^{d^n_ {14}}+e^{-d^n_{14}})(e^{d^n_{23}}+e^{-d^n_{23}})}$$ $$=\lim \frac{e^{d^n_{13}}e^{d^n_{24}}}{e^{d^n_{14}}e^{d^n_{23}}}+\lim \frac{e^{ d^n_{13}}}{e^{d^n_{14}}e^{d^n_{23}}}+\lim \frac{e^{d^n_{24}}}{e^{d^n_{14}}e^{d^n _{23}}}$$ $$=\lim e^{d^n_{13}+d^n_{24}-d^n_{14}-d^n_{23}}+\lim e^{d^n_{13}-d^n_{14}-d^n_{2 3}}+\lim e^{d^n_{24}-d^n_{14}-d^n_{23}}$$ $$=\lim e^{l(a^n)+l(b^n)-l(a^nb^n)}+\lim e^{l(a^n)-l(a^nb^n)}+\lim e^{l(b^n)-l(a ^nb^n)}$$ Since $l(a^n)+l(b^n)-l(a^nb^n)$ always exists,$l(a^n)-l(a^nb^n),l(b^n)-l(a^nb^n) $ both go to $-\infty.$ For more general argument see [Kim]. So we finally get $$\lim e^{l(a^n)+l(b^n)-l(a^nb^n)}=[x_1,x_2,x_3,x_4]$$ \end{pf} This lemma shows that marked length spectrum determines cross ratio on the limit set because every two points in the limit set can be approximated by end points of some hyperbolic isometry.By using limit argument it is clear that cross ratio of every four points in the limit set is determined by its marked length spectrum. Now we are ready to prove the theorem. \begin{th}Let $\rho:G\rightarrow Iso^{\pm}(H^n_K),\phi:G\rightarrow Iso^{\pm}(H^ m_F)$ be two nonelementary,irreducible representations having the same marked le ngth spectrum such that $n\times dim_R(K)=m \times dim_R(F)$ where $K,F$ is real ,complex,quaternion or cayley fields.Then $n=m,K=F$ and they are conjugate. \end{th} \begin{pf} After we conjugate representations we may assume that 0,and $\infty$ are in the limit sets and they are the two fixed points of hyperbolic isometries $a$ and $a'$ in $\rho$ and $\phi$ respectively.Furthermore we have $$\frac{|x_i|}{|x_j|}=|[x_i,x_j,[0,0],\infty]|^{1/2}=|[y_i,y_j,[0,0],\infty]|^{ 1/2}=\frac{|y_i|}{|y_j|}$$ So after scaling we can assume that $|x_i|=|y_i|$. Similarly $$\frac{d(x_i,x_j)}{|x_j|}=|[[0,0],x_i,\infty,x_j]|^{1/2}=|[[0,0],y_i, \infty,y_j ]|^{1/2}=\frac{d(y_i,y_j)}{|y_j|}$$ and hence $d(x_i,x_j)=d(y_i,y_j).$ Let [0,1],$[0,z]$ are the two points in the limit set of $\rho$ and $[c,d],[(t,q ),k]$ be the corresponding points in the limit set of $\phi$. Then $a$ sends $[(t,q),k]$ to $$(t',q')=e^{-2s}{\nu}^{-1}(t,q)\nu$$ \begin{eqnarray*} k'=e^{-s}\{[{\nu}^{-1}(e^{2s}+|k|^2-t,-q)^{-1}{\nu}]^{-1}\}& & \\ \{[{\nu}^{-1}((e^{2s}+|k|^2-t,-q)^{-1}(1+|k|^2-t,-q))][((1+|k|^2-t,-q)^{-1}k)M] \}& & \end{eqnarray*} where $\nu=1,M \in O(m-1),U(m-1)$ in real and complex case,$|\nu|=1,M \in Sp(m-1 ),|M|=1$ in quaternion and cayley cases. Similarly $a'$ sends $[(t,q),k]$ to $$(t',q')=e^{-2s}{\nu}^{-1}(t,q)\nu$$ \begin{eqnarray*} k'=e^{-s}\{[{\mu}^{-1}(e^{2s}+|k|^2-t,-q)^{-1}{\mu}]^{-1}\}& & \\ \{[{\mu}^{-1}((e^{2s}+|k|^2-t,-q)^{-1}(1+|k|^2-t,-q))][((1+|k|^2-t,-q)^{-1}k)N] \}& & \end{eqnarray*} Now $$d([0,z],[0,1])^4=d([(t,q),k],[c,d])^4$$ gives us that $|(t,q)|$ is a function of $z,k,c,d$.But $$d([0,z],a^n([0,1]))^4=d([(t,q),k],a'^n([c,d]))^4$$ gives that $|(t,q)|$ is a function of $e^{-ns},z,k,c,d$ for all n.This is because right side of equation has the higher power of $e^{-s}$.This is not possible unless $|(t,q)|=0$. By this way we can conclude that all the limit sets of $\phi$ corresponding to $[0,z]$ of $\rho$ are the form of $[0,k]$.Then by further conjugation we can assume that [0,1] is also the corresponding point of $\phi$ to [0,1] of $\rho$. Next we want to show that either $M=N,\nu=\mu$ or $M=\bar{N},\nu=\bar{\mu}$. This can be seen from the equation $$d(a^n([0,1]),[0,1])^4=d(a'^n([0,1]),[0,1])^4$$ for all n. Now by conjugating the representation by the map $(k \rightarrow \bar{k})$ we can assume that $M=N,\nu=\mu$. If $[0,w]$ is a point in the limit set of $\rho$ and $[0,w']$ is the corresponding point of $\phi$ then $$d(a^n([0,w]),[0,1])^4=d(a'^n([0,w']),[0,1])^4$$ gives that $w=w'$. These all show that $K=F,m=n$.So far we conjugated two representations to get that two limit sets are equal on the set $K^{m-1}=\{[0,k]:k \in R^{m-1},C^{m-1},Q^{m-1}, O\}$ and two spaces are actually the same one of the four hyperbolic spaces. For any point $x$ of the limit set of $\rho$ not in the set $K^{m-1}$ there are only two possible corresponding points $y$(either $x$ itself or reflection of $x$ along $K^{m-1}$) such that the distance between $x$ and every point in the limit set of $\rho$ lying on the set $K^{m-1}$ is equal to the distance between $y$ and corresponding point of $\phi$. By taking reflection along this set if necessary,we can assume that $x=y$. Now for any point in the limit set of $\rho$ off the set $K^{m-1} \cup x$ there is a unique corresponding point as above.So two limit sets are actually equal. This shows that we found an isometry such that after we conjugate one representation by this isometry the limit sets are equal.But this actually shows that two representations are equal.The reason is the following. Note that the end points of hyperbolic isometry $aba^{-1}$ are the images of end points of $b$ under $a$.But since every isometry is determined by images of finitely many points on $\partial{H^m_F}$ and since two representations have the same limit sets after conjugation each element should be the same. \end{pf} \begin{co}Let M,N be rank one locally symmetric manifolds of the same dimension and homotopically equivalent.If none of them have totally geodesically embedded submanifold of dimension greater than one and if they have the same marked length spectrum then they are isometric. \end{co} Since three manifolds get special attention we will write down the corollary for hyperbolic three manifolds.Here we do not need the assumption about irreducibility. \begin{co}If we have two nonelementary orientable hyperbolic three manifolds having the same marked length spectrum then they are isometric. \end{co} \begin{pf}Since every orientation preserving Mobius map is determined by the images of three distinct points whether they are on the line or not we don't have to worry about irreducibility. \end{pf} \section{Space of Representations into $SL_2(C)$} \subsection{Backgroud} Let G be a finitely generated group.Space of representations from G to $SL_2(C)$ is a set of homomorphisms from G into $SL_2(C)$.We say that two representations are equivalent if they are conjugated by some element of $SL_2(C)$.The character of a representation $\rho$ is the function $\chi_{\rho}:G \rightarrow C$ such that $\chi_{\rho}(g)=tr(\rho(g))$. If $G=\{g_i,i=1,\cdots,n;r_1=\cdots=r_k=1\}$ then space of representations $R(G)$ is a subset of $SL_2(C) \subset C^{4n}$,wich are set of all points $(\rho(g_1), \cdots,\rho(g_n))$ satisfying the relator relations.It is easy to see that $R(G)$ is an affine algebraic set in $C^{4n}$.See [CS] for details. If we have a finite volume complete hyperbolic 3-manifold then its holonomy representation is discrete ,faithful and irreducible. For each $g \in G$ we define a regular function $\tau_g$ on $R(G)$ by $\tau_g(\rho)=tr\rho(g)$. If T is a ring generated by all functions $\tau_g,g\in G$ then it is finitely generated.See [CS] for a proof. Also it is shown in [CS] that character space of an irreducible component of $R( G)$ is an affine variety. Thurston([Th] chapter 5) showed that geometric structure of compact manifold M are determined by holonomy representations of $\pi_1(M)$ near some geometric structure and its complex dimension is at least $3 \times(\#(generators)-\#(relators)-1)$. \begin{th}([Th],[CS])Let N be a compact orientable 3-manifold.Let $\rho:\pi_1(N) \rightarrow SL_2(C)$ be an irreducible representation such that for each torus component T of $\partial{N}$,$\rho(im(\pi_1(T) \rightarrow \pi_1(N)))$ not trivial. Let R be an irreducible component of $R(\pi_1(N))$ containing $\rho$.Then $X_0=character(R)$ has dimension $\geq s-3\chi(N)$,where s is the number of tori component of $\partial{N}$. \end{th} \subsection{Local smooth coordinate chart of Representation space by lengthes of finite number of loops} Let $G$ be a group and $\Re$ be the representation space of $G$ into $SL_2(C)$ ,where $\Re$ is the set of equivalence classes of homomorphisms from $G$ into $SL_2(C)$ and two representations are equivalent if they are conjugate . Suppose $\rho$ is a representation sitting smoothly in $\Re$.We want to find nice coordinate chart around $\rho$ such that each coordinate is a translation length of some element in $G$. \begin{lemma} Let $\alpha$ be an isometry in $SL_2(C)$. Then $$ |tr(\alpha)-2|+|tr(\alpha)+2|=2(e^{\frac{l(\alpha)}{2}}+e^{-\frac{l(\alpha)}{ 2}}) $$ \end{lemma} \begin{pf} If $\alpha$ is \[ \alpha=\left[ \begin{array}{cc} \lambda & 0 \\ 0 & \lambda^{-1} \end{array} \right] \] Put $x=|\lambda+\lambda^{-1}-2|,y=|\lambda+\lambda^{-1}+2|$. Then $x^2+y^2=2|\lambda|^2+2\lambda\overline {\lambda^{-1}}+2\lambda^{-1}\bar{\lambda} +2|\lambda^{-1}|^2+8$ and $$x^2+y^2+xy=4(|\lambda|^2+2+|\lambda^{-1}|^2).$$ By using $|\lambda|=e^{\frac{l(\alpha)}{2}}$ claim follows. \end{pf} Now let $\acute{G}$ be a subset of $G$ consisting of all elements whose image under $\rho$ is either hyperbolic or elliptic. Notice that if $a,b \in{\acute{G}}$ then $a^n,b^n,a^nb^n$ are all in $\acute{G}$ for large n. Also if $a\in G$ and $b\in {\acute{G}}$ then $aba^{-1}$ is in $\acute{G}$.This implies that marked length spectrum on $G'$ determines representation up to conjugacy.Choose finite set S of $G$ so that traces of those elements determines representation up to conjugation and trace map from the neighborhood to ${C^*}^S$ where $C^*=C-\{2,-2\}$ is an immersion.This is possible if the representation is not elementary.See the arguement after the example below for justification. \begin{lemma}(Vogt) Let G be a free group with three generators.Let $trX_i=x_i,trX_iX_j=y_{ij },trX_iX_jX_k=z_{ijk}$.Define $$ P=x_1y_{23}+x_2y_{13}+x_3y_{12}-x_1x_2x_3$$ $$ Q=x_1^2+x_2^2+x_3^2+y_{12}^2+y_{13}^2+y_{23}^2+y_{12}y_{13}y_{23}-x_1x_2y_{12 }\\-x_1x_3y_{13}-x_2x_3y_{23}-4$$ $$\Delta(X_1,X_2,X_3)=P^2-4Q$$ Then $z_{123}$ and $z_{213}$ are roots of the quadratic equation for $z$: $$z^2-Pz+Q=0$$Given prescribed values $x_i,y_{ij}$ there exists only one conjugacy class if and only if $\Delta(X_1,X_2,X_3)=0.$ \end{lemma} See [Ma 1] for the proof. {\bf{Example}}.Let $G$ be a free group with three generators.I want to show that trace map is not an immersion sometimes.Let \[ X=\left( \begin{array}{cc} \alpha & 0 \\ 0 & {\alpha}^{-1} \end{array} \right) \] and \[ Y=\left( \begin{array}{cc} \beta & 0 \\ 0 & {\beta}^{-1} \end{array} \right) \] Let $Z$ be a third element which does not commute with any of two elements and $\Delta(X,Y,Z)\neq 0$. Since $G$ is free neighborhood of this representation is a regular neighborhood of $SL_2(C) \times SL_2(C) \times SL_2(C) $ and obviously sitting smoothly in $\Re$.Let tr be a map from this neighborhood to $C^7$ such that $$tr(\rho)=(trX,trY ,trZ,trXY,trXZ,trYZ,trXYZ)$$Then elementary calculation shows that \begin{eqnarray*} dtr_{(X,Y,Z)}(\xi_1,\xi_2,\xi_3)&=&(tr(X\xi_1),tr(Y\xi_2),tr(Z\xi_3),tr(X\xi_2) \\ & & \mbox{} +tr(Y\xi_1),tr(X\xi_3)+tr(Z\xi_1),tr(Y\xi_3)+tr(Z\xi_2),...) \end{eqnarray*} where $\xi_i \in sl_2(C).$ Then dimension of kernel at $(X,Y,Z)$ is 4.To see this note that there are two degrees of freedom from $tr(X\xi_1)$ and so on ,so there are six degrees of freedom from $\xi_1,\xi_2,\xi_3$. $tr(X\xi_1), tr(Y\xi_2)$ equal to zero implies $tr(X\xi_2)+tr(Y\xi_1)$ is zero since $X$ and $Y$ commute.But it should satisfy $tr(X\xi_3)+tr(Z\xi_1)=0$ and $tr(Y\xi_3)+tr(Z\xi_2)=0$ so it takes out two degrees of freedom. If $\xi_1,\xi_2,\xi_3$ satisfy $tr(X\xi_1)=tr(Y\xi_2)=tr(Z\xi_3)=tr(X\xi_2)+tr(Y\xi_1) =tr(X\xi_3)+tr(Z\xi_1)=tr(Y\xi_3)+tr(Z\xi_2)=0$ then the last term vanishes automatically.Here is the reseason. If $X_i(t)$ are curves passing throug $X,Y,Z$ respectively and with tangent vectors $\xi_1,\xi_2,\xi_3$ set $$tr(X_i(t))=x_i(t) ,trX_i(t)X _j(t)=y_{ij}(t),trX_1(t)X_2(t)X_3(t)=z(t).$$ By lemma 3 $z(t)^2 -P(t)z(t)+Q(t)=0.$If we differentiate this equation at 0 we get $z'(0)(2z(0)-P(0)=0$ by using $P'(0)=Q'(0)=0$.But $2z(0)-P(0)\neq 0$ since $\Delta$ is not equal to zero.So tr is not an embedding at this point. This example has an obvious geometric meaning.For $tr(X\xi_1)$ to be zero, diagonal elements of $\xi_1$ should be equal to zero.This implies that we deform $X$ by conjugation. The same is true for $Y$.Since $X,Y$ have the same axis, conjugating $X$ by those elements conjugating $Y$ does not change trace of $X$ so infinitesimal change of $tr(X)$ is equal to zero. Same argument holds for $Y$. More explicitly let \[ g_t=\left( \begin{array}{cc} x_1(t) & x_2(t) \\ x_3(t) & x_4(t) \end{array} \right) \] and \[ h_t=\left( \begin{array}{cc} y_1(t) & y_2(t) \\ y_3(t) & y_4(t) \end{array} \right) \] such that $g_0=h_0=I$ and $x_2'(0)=-y_2'(0)=x_3'(0)=-y_3'(0)$. Let \[ Z=\left( \begin{array}{cc} a & b \\ b & d \end{array} \right) \] for simplicity. Set $X_t=g_tX{g_t}^{-1},Y_t=h_tY{h_t}^{-1},Z_t=g_tZ{g_t}^{-1}$. Then ${\xi(t)}=(X_t,Y_t,Z_t)$ is a 1-parameter family of defermation which is not a conjugation defermation.For \[ X'(0)= \left( \begin{array}{cc} 0 & -x_2'\alpha+x_2'{\alpha}^{-1}\\ x_3'\alpha-x_3'{\alpha}^{-1} & 0 \end{array} \right) \] and similarly for $Y'(0)$ there is no conjugation deformation giving tangent vectors $X'(0)$ and $Y'(0)$.It is easy to see that $dtr_{(X,Y,Z)}(\xi'(0))=0$ .For if we calculate $\frac{dtr(Y_tZ_t)}{dt}$ at $t=0$,it is equal to $\beta(x_2'(0)b-x_3'(0)b)+ b(-y_2'(0)\beta+x_2'(0){\beta}^{-1})+b(y_3'(0)\beta-y _3'(0){\beta}^{-1})+{\beta}^{-1}(-x_2'(0)b+x_3'(0)b)$. If we look at above example carefully,it shows if no two elements commute then trace map is an immersion,which implies that every tangent vector in the kernel of $dtr$ comes from some conjugation deformation. Observe the following calculation. \[ g_t=\left( \begin{array}{cc} x_1(t) & x_2(t) \\ x_3(t) & x_4(t) \end{array} \right) \] \[X= \left( \begin{array}{cc} a & b \\ c & d \end{array} \right) \] Then conjugation of $X$ by $g_t$ give tangent vector at $X$ \[ \left( \begin{array}{cc} x_2'c-x_3'b & -x_2'a+2x_1'b+x_2'd \\ x_3'a+2x_4'c-x_3'd & -x_2'c+x_3'b \end{array} \right) \] Notice that two entries are linearly independent.So if $g_t,h_t$ give the same infinitesimal deformation on two different element which do not commute then they give the same infinitesimal deformation on every element. By using this fact we can show that if the representation is not elementary then we can show that trace map from a neighborhood of a representation $\rho$ sitting smoothly in representation space to $C^N$ for large enough N is a smooth immersion. To see this if $S=\{X_1,\cdots,X_n\}$ is the finite set containing all generators of the group such that its traces determine representation uniquely up to conjugacy(which is possible by [CS]) and their images under $\rho$ are all hyperbolic isometries.By abusing the notation we will identify $X_i$ with $\rho(X_i)$. This can be seen that if $S$ is such a set possibly contaning parabolic or elliptic elements then choose a hyperbolic element and multiply those elements which are not hyperbolic in $S$ by this element and its inverse to make them hyperbolic and then throw in this element.Then using the formula $tr(XY)+tr(XY^{-1})=tr(X)tr(Y)$ we see that all the traces of old elements can be obtained by new elements.This is a new set consisting entirely of hyperbolic elements. Since representation is not elementary there is $X_k$ which does not commute with $X_1$.Choose any $X_i,i\neq1,k$.If it commutes with $X_1$ replace $X_1,X_i, X_k $ by $X_1X_i,X_1X_k,X_k$.It is easy to check that no two of three does commute. By renaming them we get $(X_1,\cdots,X_n)$ such that no two of $X_1,X_2,X_3$ commute and $X_i,i\neq 1,2,3$ commutes with at most one of three $X_1,X_2,X_3$. If $(\xi_1,\cdots,\xi_n)$ is in the kernel of the differential of the trace map, there is $g_t$ conjugating $X_1,X_2,X_3$, giving tagent vectors $\xi_1,\xi_ 2,\xi_3$ because no two of three commute.Similary $h_t$ conjugating $X_1,X_2,X_4 $ giving $\xi_1,\xi_2,\xi_4$ since no two of three commute.But $h_t$ can be replaced by $g_t$ since they give the same infinitesimal deformation on $X_1,X_2$ which do not commute.In this way we can see that this tangent vector comes from conjugation deformation. By lemma 2 the map $f$ from ${C^*}^S$ to $R^{\acute{G}}$ sending each component $\alpha$ of an element in ${C^*}^ S$ to $2(e^{\frac{l(\alpha)}{2}}+e^{-\frac{l(\alpha)}{ 2}})$ is $C^{\infty}$ since there is no parabolic element in $\acute G$ and norm function is smooth at $C^*$.Furthermore by adding more elements in $S$ we can make this function an immersion.To see this consider the map $g$ from $R^2$ to $R$ such that $g(x,y)=|(x,y)-2|+|(x,y)+2|$ as in the above lemma.It is easy to see that if $(x,y)\neq 2,-2$ then dimension of the kernel of $dg$ is one.Suppose $dg(\xi)=0$ at $\alpha$. Then find an element $\beta$ so that trace($\beta$) is a polynomial in $\alpha$ and other elements in $S$.Choosing $\beta$ carefully it is possible to ensure that $dg(\xi) \neq 0$ at $\beta$.So adding $\beta$ to $S$ for each such an $\alpha$ we can show the map is an immersion. \begin{Prop} Let $\rho$ be an nonelementary representation sitting smoothly inside the representation space.Then neighborhood of $\rho$,denoted by $N(\rho)$, is parametrized by translation lengthes of finitely many hyperbolic elements in $G$. \end{Prop} \begin{pf} So far we got $$f\circ tr:N(\rho)\rightarrow {C^*}^S \rightarrow R^{\acute{G}}$$ is an immersion.But since marked length spectrum on $R^{\acute{G}}$ determines representation it is one to one also. Since dimension of representation space is finite around $\rho$ we can find finite subspace $R^N$ such that projection of the image of the neighborhood to this subspace is injective. \end{pf} Specially if the representation is an holonomy of geometrically finite 3-manifold then it is automatically sitting smoothly in representation space because Teichmuller space is open in representation space and it satisfies all the necessary condition for above considerations,so small neighborhood of this representation is parametrized smoothly by lengthes of finitely many geodesics. $$ ACKNOWLEDGEMENT$$ I wish to thank all whom I talked with,among them I want to specially thank A.Casson as my adviser,C.McMullen,J.P.Otal,F.Bonahon,M.Bourdon,\\ G.Besson for their useful comments.
\section{Introduction} \label{intro} Since the discovery of the dusty disk around $\beta\:$Pictoris\ (Aumann 1984, Smith \& Terrile 1984), it has been the subject of extensive investigations with all possible observational techniques: visible and IR imagery (Smith \& Terrile 1987, Paresce \& Burrows 1987, Lecavelier des Etangs et al. 1993, Golimowski et al. 1993, Lagage \& Pantin 1994), spectroscopy (Vidal-Madjar et al. 1994, Vidal-Madjar \& Ferlet 1995 and references therein) and recently photometric variations (Lecavelier des Etangs et al. 1995). In particular, the stellar light scattered by dust particles has been well observed by several authors for many years to follow a power law distribution with a slope in the range -3.6 (Artymowicz et al. 1989, Lecavelier des Etangs et al. 1993) -4.3 (Smith \& Terrile 1984). However, a convincing explanation for this scattered light distribution is still needed. The dust distribution in the $\beta\:$Pictoris\ disk may be a crucial issue for understanding its origin. We present here a precise dynamical model which will be compared with the observations for the first time in a quantitative manner. Another key point revealed by the spectroscopic studies of the gaseous counterpart of the $\beta\:$Pictoris\ disk is the existence of orbiting kilometer size bodies which sometimes fall onto the star, giving rise to the observed spectroscopic redshifted signatures due to ejected gas, and undoubtedly to dust particles. Already in 1984, Weissman (1984) questioned whether the material around Vega-like stars is of cometary or asteroidal origin and suggested that particles could be continually supplied by sublimation or from collisions between larger bodies. The time scale evaluation over which dust particles are eliminated by the Poynting-Robertson effect or collisions (Backman \& Paresce 1993) leads to durations much shorter than the estimated age of $\beta\:$Pictoris\ ($\sim 2\cdot 10^{8}$ years according to Paresce, 1991). This simple comparison suggests convincingly that the $\beta\:$Pictoris\ disk is not a remnant of planetary formation but on the contrary must be continually replenished by secondary sources like evaporation or collision of small bodies. This seems in fact a very natural hypothesis (see for example Zuckerman \& Becklin 1993): the presence of small bodies is a consequence of planetary systems formation (Lissauer 1993). Furthermore, CO absorptions have been recently observed towards $\beta\:$Pictoris\ (Vidal-Madjar et al. 1994) with the Hubble Space Telescope. The very presence of CO may also require a permanent source provided by evaporation of comet-like bodies. Here we present a new argument: This replenishment is able to reproduce the main characteristics of the $\beta\:$Pictoris\ dust disk, and in particular the spatial distribution of dust at large distances. In detail, it gives possible explanations for the following unexplained issues: \begin{enumerate} \item The gradient of the scattered light follows a relatively well-known but unexplained power law (e.g. Golimowski et al. 1993). \item The distribution at large distances is obviously not axisymmetric (Smith \& Terrile 1987, Kallas \& Jewitt 1995). \item The central part of the disk is relatively clear of dust (Backman et al. 1992, Lagage \& Pantin 1994). \item The disk seems to be a ``wedge'' disk: the thickness increases with radius (Backman \& Paresce 1993, Lecavelier des Etangs et al. 1993). \item The slope of the scattered light distribution changes abruptly at about 100~AU from the star. If confirmed, this fact remains unexplained (Artymowicz et al. 1990, Golimowski et al. 1993). \end{enumerate} Furthermore, the possible explanation of these observational facts could give new understanding to the following questions: \begin{itemize} \item a) What is the mass of the dust disk, since when extrapolated towards infinite distances the disk mass diverges (Artymowicz et al. 1989)? \item b) Are there connections between the dust and the gas disks ? \item c) Can the asymmetry in the observed dust disk be connected with the asymmetry in the longitude of periastron of the comets in the Falling-Evaporating-Bodies (FEB) model proposed to explain the redshifted spectroscopic events (e.g. Beust et al. 1991)? \end{itemize} Therefore, we will consider here a model in which the disk is replenished by a group of kilometer-size bodies (Section~\ref{model}). Numerical calculations of the spatial distribution of dust are given in Section~\ref{numerical}. We shall discuss in Section~\ref{bp disk} the $\beta\:$Pictoris\ disk in such a scenario and summarize the results in Section~\ref{conclusion}. \section{The model} \label{model} \subsection{Basic concept} We assume that the dust is produced by cometary or asteroid-like bodies which create particles through mutual collisions or evaporation processes. The main assumption is thus that these particles have small initial velocities relative to the parent bodies: less than 1 km/s like in cometary production (Gombosi et al. 1985, Sekanina 1987) or in "Chiron burst" (Luu \& Jewitt 1990). Thus, these velocities can be considered as negligible in comparison to the orbital velocities. In our model, this relative velocity is fixed at zero. However, as soon as a dust particle is ejected, it is perturbed by the radiation pressure; its orbit is different from the parent body one and tangent at the point where it was injected (Burns et al. 1979). For example, if a parent body in a circular orbit with a semi-major axis $a_0$ produces a particle with a ratio $\beta $ of the radiation force to the gravitational force, the particle orbit has a semi-major axis $a_{\beta}=a_0(1-\beta )/(1-2\beta )$ and an eccentricity $e_{\beta}=\beta /(1-\beta )$. The periastron is $a_{\beta}(1-e_{\beta})=a_0$ and the apoastron $a_{\beta}(1+e_{\beta})=a_0/(1-2\beta )$. Therefore, particles can be observed at distances from the central star much larger than the apoastron of the parent body. Thus, local perturbations on the distribution of the parent bodies like asymmetries, could produce observable signatures on the dust distribution at very large distances. \subsection{Particle size distribution} \label{size} Furthermore, because the parent bodies should not produce single sized particles, we introduce in the calculation a size distribution. For particles with radius $s\ge 1\mu$, $\beta$ is correlated with the size of the particle by $\beta\sim s^{-1}$ (Artymowicz 1988). We can assume a power law size distribution as in Solar System cometary dust (Lien 1990 and references therein), in collisionally replenished dust (Greenberg \& Nolan 1989), or in interplanetary medium dust (Le Sergeant \& Lamy 1980). Thus, if we have a size distribution $dn\propto s^{q}ds$, then we have $dn\propto \beta^{K}d\beta$ with $K=-q-2$. \subsection{Analytical consideration} \label{analytical} \subsubsection{One point production} First, let us consider a parent body in a circular orbit which generates a set of particles at a given point. Then the particles with $\beta >0.5$ have hyperbolic orbits and are ejected from the system towards the interstellar medium. The other particles with $\beta <0.5$ follow orbits within a parabola which is the orbit of the particles with $\beta=0.5$. One can evaluate the surface density of particles in the asymptotic direction of the parabola: for a given particle the true anomaly $\theta$ follows the distribution law \begin{center} \begin{equation} P_{\theta}(\theta )d\theta =\frac{(1-e)^{3/2}} {2\pi (1+e\cos \theta)^2} d\theta \end{equation} \end{center} In the asymptotic direction $\theta=\pi$, $P_{\theta}(\pi)\propto (1-e)^{-1/2} \propto r^{1/2} (1+a_0/r)^{1/2}$. The probability that the particle apoastron is between $r$ and $r+dr$ in the same direction is $P_r(r)= \beta^K (d\beta /dr) dr$. Since $\beta=(1-a_0/r)/2$, $P_r(r)\propto (1-a_0/r)^K r^{-2} dr$. Thus, taking into account the scattering cross section $\sigma\propto \beta^{-2}$ we obtain the surface density normal to the plane of the disk $\sigma \cdot n_s$: \begin{center} \begin{equation} \sigma \cdot 2\pi r\cdot n_s(r)\propto r^{-3/2}(1-a_0/r)^K (1+a_0/r)^{1/2} \end{equation} \end{center} Thus $\sigma \cdot n_s(r)_{r \rightarrow \infty}\propto r^{-2.5}$. \subsubsection{Parent bodies in circular orbits.} \label{circular orbits} However, dust production can take place at every point in the orbit of the parent body. Here, we restrict ourselves to a parent body in circular orbit, thus, producing an axisymmetrical disk. Each particle has a probability $f_{\beta}(r)$ to be at a distance $r$: $f_{\beta}(r)\propto ((1-a_0/r)(2\beta -1 +a_0/r))^{-1/2}$. The surface density normal to the plane of the disk is then: \begin{center} \begin{equation} \sigma \cdot n_s(r)\propto \int_{(r-a_0)/2r}^{1/2} \frac{\beta^{K-2} f_{\beta}(r) d\beta } {2 \pi r\int_{a_0}^{a_0/(1-2\beta)} f_{\beta}(r_1)dr_1} \label{circular} \end{equation} \end{center} This surface density follows an $r^{-3}$ law and is plotted in Fig.~\ref{fig1}. Moreover, the inclination of the particles is the same as the inclination of the parent bodies. Therefore, the vertical distribution of the particles depends on the distribution of the inclination of the parent bodies. In the present model the thickness of the disk is increasing with radius and the volume density ($n_v$) is proportional to the surface density ($n_s$) divided by the distance, i.e. $n_v(r)\propto r^{-4}$. It has to be noted that this conclusion is valid at distances larger than the distance of the farthest parent bodies. Within this zone the opening angle must look smaller. {}From the $n_v(r)\propto r^{-4}$ law and Nakano's (1990) conclusion concerning the connection between the volume density and the scattering light distribution ($F(r)\propto n_v(r)/r $), one can conclude that a belt of parent bodies in circular orbits should produce a "wedge" disk with a scattered light distribution $F(r)\propto r^{-5}$. \begin{figure}[tbp] \ifthisfig \psfig{file=Disk.ps,height=\columnwidth,rheight=8.5truecm} \else\picplace{6truecm}\ifnofig\thisfigfalse\else\thisfigtrue\fi\fi \caption[]{Plot of the surface density distribution calculated from Eq.~\ref{circular}. The abrupt density decrease at $r\sim a_0$ is simply due to the assumed single distance between the central star and all the parent bodies. It is thus an artificial consequence of an oversimplified hypothesis.\\ For comparison, the hatched zone represents the distribution range of slopes deduced by different observers around $\beta\:$Pictoris.} \label{fig1} \end{figure} \section{Numerical results} \label{numerical} \subsection{The Monte-Carlo simulation} The analytical model of the previous section can only be solved for a limited number of parameters and parent body distributions. We have chosen to evaluate the dust distribution by a Monte-Carlo method. We shall consider a given family of parent bodies, a dust production law and a particle size distribution, and we shall integrate the distribution in the following way: For each of the $N$ random particles, we first randomly select a parent body in the family and its position in its orbit; then we randomly apply the particle size distribution and we calculate the position $r_i$ of the particle $i$ after a time $t$ where $t$ is randomly chosen between $0$ and $t_{max}$. Finally, we evaluate $F(r)$ the surface brightness along the midplane of the edge-on disk projected on the sky at the projected distance $r$. If $F_i(r)$ is the contribution of the particle $i$, from Eq.~2 of Nakano (1990), we have \begin{center} \begin{equation} F_i(r)\propto r^{-1}\int _0 ^{\pi}n_i(\Lambda)\sigma_i(\theta)d\theta \end{equation} \end{center} where $\Lambda$ is the distance to the central star and $\sigma_i(\theta)$ is the scattering phase function. Nakano (1990) has already shown that, except if the dust is much more steeply forward-scattering than in the Solar System, the isotropic scattering assumption does not change the observed slope of the scattered light distribution, thus $\sigma_i(\theta)=\beta _i^{-2}$ with $\sin \theta = r/\Lambda$. The volume density is given by $n_i(\Lambda) = \delta(\Lambda-r_i)\cdot p_{i} /r_i$ where $p_i$ is the production rate of dust on the parent bodies, then $F_i(r)\propto r^{-1}\int \delta(\Lambda-r_i) p_i {\rm H}(\Lambda-r)r d\Lambda / (r_i \beta _i^{2} \Lambda \sqrt{\Lambda^2-r^2})$ where H is the Heavyside function. Finally, the surface brightness at a given projected distance $r$ from the star is estimated by \begin{center} \begin{equation} F(r)=\sum_{i=1}^{N}F_i(r) \propto \sum_{i=1}^{N}\frac{ {\rm H}(r_i-r)\cdot p_{i} } {r_{i}^{2} \sqrt{r_i^2-r^2}\cdot \beta _{i}^{2}} \end{equation} \end{center} \subsection{Results} Results of runs with different initial conditions on the family of parent bodies are summarized in Tables~\ref{resultat_1} and \ref{resultat_2}. The production rates per annulus of fixed width are given for each run. Several configurations have been tried. The first one corresponds to the analytical model of Sect.~\ref{circular orbits}, we find again $\alpha=5.0$, which validates the Monte-Carlo model. In runs \#2 and \#3, we have tried different distances and evolutionary times which show that the Poynting-Robertson effect does not change the results: the accuracy for $\alpha$ is about 0.1. From \#4 to \#11, the parent bodies are in eccentric orbits with eccentricities uniformly distributed between 0 and 0.5 or between 0 and 1; it seems that bodies with eccentricities larger than 0.5 do not contribute much to the distribution since a large fraction of ejected material is then on hyperbolic orbits. In runs \#4 to \#7, we have taken parent bodies with a peculiar longitude of periastron ($\omega$), namely 155$^o$. This value has been selected according to the Beust et al. (1991) modelisation of the Falling-Evaporating-Bodies, the so-called FEB scenario needed to reproduce the spectroscopic redshifted signatures. In the four last runs of Table~\ref{resultat_1}, $\omega$ was uniformly distributed between 0$^o$ and 360$^o$. {}From all the results in Table~\ref{resultat_1}, it can be seen that the slope $\alpha$ is always greater than or equal to 5.0. In Table~\ref{resultat_2}, we test other parent body distributions in which the dust observed at large distances still is produced close to the star. Incidentally, another solution would be to assume that the disk is produced by a distribution of asteroids up to 1000~AU from the star. However this solution does not allow us to explain the observed asymmetries. Runs \# 12 to \# 15 take into account the Epstein gas drag with a gas density $\rho_{gas}=\rho _0 (100\ {\rm AU}/r)$ and temperature $T_{gas}=20$~K. Runs \# 16 and \# 17 are with dust size distribution characterized by $K=5-10$. \begin{table*} \caption{Slope $\alpha$ of the gradient of brightness in disks with different conditions for the parents bodies. In all these runs we assumed N=10000 and K=1.5. Poynting-Robertson drag is taken into account but there is no gas drag. For runs \#~4 to \#~7, the two slopes are for the both sides of the asymmetrical disk seen from the Earth.} \label{resultat_1} \begin{tabular}{|c|cccc|c|c|} \hline Run&&Parent Bodies Parameters &&&$t_{max}$ &Result: $\alpha$\\ $\#$ & semi-major axis (AU) & Eccentricity & $\omega$ & Production rate & (years)&$F(r)\propto r^{-\alpha}$ \\ \hline 1 & 20 & 0. & - & $p=const$ & 10000 & 5.0 \\ 2 & 2 & 0. & - & $p=const$ & 10000 & 4.9 \\ 3 & 20 & 0. & - & $p=const$ & 100000 & 5.0 \\ 4 & 20-30 & 0.-0.5 & 155$^o$ & $p=const$ & 10000 & 5.2-5.6 \\ 5 & 20-30 & 0.-1. & 155$^o$ & $p=const$ & 10000 & 5.2-5.7 \\ 6 & 20-30 & 0.-0.5 & 155$^o$ & $p\propto r^{-3}$ & 10000 & 5.3-5.5 \\ 7 & 20-30 & 0.-1. & 155$^o$ & $p\propto r^{-3}$ & 10000 & 5.3-5.6 \\ 8 & 20-30 & 0.-0.5 & 0-360$^o$ & $p=const$ & 10000 & 5.2 \\ 9 & 20-30 & 0.-1. & 0-360$^o$ & $p=const$ & 10000 & 5.3 \\ 10 & 20-30 & 0.-0.5 & 0-360$^o$ & $p\propto r^{-3}$ & 10000 & 5.2 \\ 11 & 20-30 & 0.-1. & 0-360$^o$ & $p\propto r^{-3}$ & 10000 & 5.4 \\ \hline \end{tabular} \end{table*} \begin{table*} \caption{Same as the previous table, with different $K$; gas drag is taken into account. The parent bodies have semi-major axis $a=$20~AU and eccentricity $e=0$. $t_{max}$=10000. For the runs \# 16 the slope gradually changes from $\alpha=3.7$ at 20 AU to $\alpha=4.7$ at 800 AU; and for run \# 17 from 2.9 to 4.4 respectively.} \label{resultat_2} \begin{center} \begin{tabular}{|c|c|c|c|} \hline Run & $\beta$ distribution:K& gas density at 100 AU & Result: $\alpha$ \\ $\#$ & $dn=\beta ^{K}d\beta$ & (cm$^{-3}$) & $F(r)\propto r^{-\alpha}$ \\ \hline 12 & 1.5 & 100. & 5.0 \\ 13 & 1.5 & 1000. & 5.1 \\ 14 & 1.5 & 10000. & 4.6 \\ 15 & 1.5 & 100000. & 3.8 \\ 16 & 5. & 0. & 3.7-4.7 \\ 17 & 10. & 0. & 2.9-4.4 \\ \hline \end{tabular} \end{center} \end{table*} \section{The $\beta\:$Pictoris\ disk} \label{bp disk} \subsection{Orbiting-Evaporating-Bodies (OEB). Towards a solution?} The observed distribution in the external part of the $\beta\:$Pictoris\ disk ($\alpha \le 4.3$) is less steep than all calculated slopes in Tables~\ref{resultat_1}. It has to be noted that a solution compatible with the observed slope could be obtained with a gas density $\rho_0 \sim 10^{5}$~cm$^{-3}$ at 100~AU. However, this is excluded by the upper limit deduced from spectroscopic or HI observations (Vidal-Madjar et al. 1986, Freudling et al. 1995). Runs \#16 and \#17 are also compatible with the dust distribution in $\beta\:$Pictoris\ (Fig.~\ref{16}). But they require dust size distributions ($K=5-10$) very different from the expected ones if they are produced through collisions between parent bodies (Fujiwara 1979 and references therein). However they show us that we need a relatively larger amount of smaller particles which have larger apoastrons. In the evaporation process of large bodies (like Chiron), there is an upper limit $s_{max}$ to the size of the ejected particles extracted by gas (Luu \& Jewitt 1990). Following Cowan \& A'Hearn (1982) we have calculated this upper limit and the evaporation rate $Z$ in molecules per second per unit area for CO and CO$_2$ as a function of the distance to $\beta\:$Pictoris, assuming $L_{\beta Pic}=6$L$_{\odot}$ and a typical parent body radius $R_{body}=100$~km (Figs.~\ref{Z} and \ref{smax}). This upper limit is inversely proportional to the parent body size $R_{body}$. We have introduced this effect in our model. The results are summarized in Table~\ref{resultat_co2} for CO$_2$; they would be similar for CO if the parent bodies were at larger distances. Here, we have taken the production rate $p\propto Z\cdot R_{body}^2$ and a parent bodies size distribution $dn=R_{body}^{-\gamma}dR_{body}$ where $\gamma=3.5$ (uncertainties deduced from observations in the Solar System with $\gamma$ between 3.2 (Whipple 1975, Hughes \& Daniels 1982) and 3.8 (Fern\'andez 1982) do not change the present results). In runs \# 101 to \# 103, we take a wide belt of parent bodies from 15 to 30 AU, without correlation between their radius $R_{body}$ and their distance from the central star. We obtain $\alpha \sim 4.8-4.9$; the effect of cutoff in particle size is not efficient enough to explain the observed slope $\alpha \sim 3.6-4.3$. However, if the evaporating bodies are coming from larger distances and their orbital parameters diffuse from a Kuiper belt-like zone towards a planetary-like zone, the smaller parent bodies lost their volatile material very early at larger distances whereas the larger bodies can evaporate downward at smaller distances. This scenario can be modelized in the following way: a parent body with a characteristic distance (which can be the periastron or the semi-major axis) $r(t)=r_0-\dot{r}t$ is considered as evaporating at $r=r(T)$ only if the total mass of the previously evaporated gas is smaller than the available mass of volatile material. \begin{center} \begin{equation} \label{dead comet} \int_{0}^{T} \mu Z(r(t))4\pi R_{body}^{2}\lambda dt < \frac{4\pi }{3}\rho \xi R_{body}^3 \end{equation} \end{center} where $\lambda$ is the percentage of active surface of the body, $\xi$ the relative mass of CO$_2$ in the parent body, $\rho$ its density and $\mu$ the CO$_2$ molecular weight. Eq.~\ref{dead comet} can be reduced to $R_{body}\ge R_{ref}\int_{r}^{\infty}Z(l)dl/\int_{r_{ref}}^{\infty}Z(l)dl$ where $R_{ref}$ is the reference size of the smallest body which can evaporate at $r=r_{ref}$. For CO$_{2}$ evaporation and $r_{ref}=20$~AU, one obtain \begin{center} \begin{equation} \dot{r}=6\cdot 10^{-7} \frac{\lambda}{\xi \rho} \left( \frac{10 {\rm km}} {R_{ref}} \right){\rm AU\ year}^{-1} \end{equation} \end{center} In runs \# 104 to \# 115, only those with $R_{ref}\sim 30-40$~km give the observed slope range. Moreover, due to the particle size cutoff, $\beta \geq 0.4$ there is an abrupt break in the slope at $r_{break}\approx a_0/(1-2\beta_{min}) \approx 100\ {\rm AU}$ (see Fig.~\ref{106} for run \#106). For CO, one can obtain $\dot{r}=6\cdot 10^{-6} \lambda /(\xi \rho) {\rm AU\ year}^{-1}$, and there would be also an abrupt break in the slope since CO begins to evaporate from about 100-150 AU. \begin{figure}[tbp] \ifthisfig \psfig{file=s12.ps,height=\columnwidth,rheight=8.5truecm} \else\picplace{6truecm}\ifnofig\thisfigfalse\else\thisfigtrue\fi\fi \caption[]{Plot of the surface brightness (arbitrary unit) of the equatorial plane of an edge-on disk seen from the Earth calculated in run \# 16. The slope varies from -3.7 to -4.7. For comparison, the dashed lines represent the power laws $r^{-3.7}$ and $r^{-4.7}$.} \label{16} \end{figure} \begin{figure}[tbp] \ifthisfig \psfig{file=Zco.ps,height=\columnwidth,rheight=8.5truecm} \else\picplace{6truecm}\ifnofig\thisfigfalse\else\thisfigtrue\fi\fi \caption[]{Plot of the evaporation rate $Z$ as a function of the distance from $\beta\:$Pictoris. The thick lines are for CO$_2$ and the thin ones for CO. The evaporation is calculated for a steady-state energy balance and depends on the albedo $A_v$ of the parent body. The short-dashed lines are for $A_v$=0.1, the long-dashed are for $A_v$=0.8 and the solid line are for $A_v$=0.5. In all the simulations we take $A_v$=0.5.} \label{Z} \end{figure} \begin{figure}[tbp] \ifthisfig \psfig{file=Amax.ps,height=\columnwidth,rheight=8.5truecm} \else\picplace{6truecm}\ifnofig\thisfigfalse\else\thisfigtrue\fi\fi \caption[]{Plot of the radius of the largest particle which can be ejected from the surface of a parent body with radius $R_{body}$=100~km and density $\rho_{body}=1$ as a function of $\beta\:$Pictoris\ distance. The thick lines are for CO$_2$ and the thin ones for CO. The different albedos $A_v$ are the same as in Fig.~3.} \label{smax} \end{figure} \subsection{Discussion} We saw in the previous section that the observed gradient in the $\beta\:$Pictoris\ disk can be explained by the present model only if there is a larger amount of small particles. Now, there is an upper limit in the size of particles ejected by slowly evaporating bodies due to a balance between gas drag and gravitation. In fact, we obtain dust disks very similar to the $\beta\:$Pictoris\ one by assuming evaporation of kilometer size bodies which are diffusing from a Kuiper belt-like zone inward an evaporation zone. The time scale $\dot{r}\sim 10^{-7}$~AU~year$^{-1}$ in the diffusion of comet orbital elements, which is necessarily to explain $\beta\:$Pictoris's characteristics, is consistent with the few AU in 10 Myr deduced by Torbett \& Smoluchowski (1990) and Levison's (1991) results on the Kuiper belt dynamical behavior between 30 and 40 AU around the Sun. The analysis of the evolution of the orbital elements of the parent bodies is outside the scope of this paper; moreover, the motion of such Kuiper belt-like bodies under the influence of a planetary system is probably chaotic (Scholl 1979, Torbett \& Smoluchowski 1990) and very dependent on this planetary system. The present model merely shows us that the dust distribution may be an important clue and may be explained by the present model if there is a long time scale perturbation of parent bodies by a planetary system which is compatible with the one predicted for the Solar System. Another evolution process for heating the bodies consists in the increase of the star luminosity ($L$) during its evolution along the main sequence. For a star of 1.6~M$_{\odot}$, we have $dL/Ldt=2\cdot 10^{-10}$~year$^{-1}$, if $\tilde{r}$ is the distance of equal luminosity, this corresponds to $\dot{\tilde{r}}=\tilde{r}/2L\cdot dL/dt$. This solution would have the advantage of being "planetary system" independent. However, if the primary volatile is CO, $\dot{\tilde{r}}=10^{-8}$~AU~year$^{-1}$; even if the volatile is CO$_2$, $\dot{\tilde{r}}=2.5\cdot 10^{-9}$~AU~year$^{-1}$. The time scales do not seem to be adequate. This mechanism would be efficient only if we could have bodies with very large quantities of volatiles and very small fractions of active surface. The assumption that CO or CO$_2$ can be primary volatiles is not a problem since around an AV star they can evaporate at very large distances where the presence of young objects without a lag deposit of nonvolatile forming a crust is likely. On the contrary in the Solar System, we observe bright comets only when H$_2$O can evaporate, simply because CO or CO$_2$ evaporate only inside planetary regions where, for dynamical reasons, the presence of young objects is unlikely: Chiron seems to be an exception but provides a useful analogy with what may happen around $\beta\:$Pictoris. As we shall see below, this model is also able to give answers to other important issues defined in Section~\ref{intro}. \begin{table*} \caption{Same as the previous tables, for particles ejected by evaporating gas. Here we assume that the gas is CO$_2$ and $p\propto Z\cdot R_{body}^2$. $R_{ref}$ is the radius of the largest dead comets at $r_{ref}$=20~AU. For runs \# 105 to \# 115 there is a slope break at $r=r_{break}$, and both inward and outward slopes $\alpha _{in}$ and $\alpha _{out}$ are given. For runs \# 113 to \# 115 the two given slopes $\alpha _{in}$ are for each side of the disk. We always took $t_{max}$=10000 years.} \label{resultat_co2} \begin{tabular}{|c|ccc|c||ccc|c|} \hline Run && Parent Bodies Parameters && $R_{ref}$& &Result: & $F(r)\propto r^{-\alpha}$ & Brightness\\ &&&& & $r_{break}$ & $\alpha _{in}$ & $\alpha_{out}$ & ratio\\ $\#$ & Periastron (AU) & Eccentricity & $\omega$ & (km) &(AU)&&&\\ \hline 101 & 20-30 & 0. & - & - & - & - & 4.8 & - \\ 102 & 20-30 & 0.-0.5 & 0-360$^o$ & - & - & - & 4.9 & - \\ 103 & 20-30 & 0.-0.5 & 155$^o$ & - & - & - & 4.9 & - \\ 104 & 15-30 & 0. & - & 10. & - & - & 4.6 & - \\ 105 & 15-30 & 0. & - & 20. & 50 & 2.9 & 4.5 & - \\ 106 & 15-30 & 0. & - & 30. & 100 & 2.7 & 4.3 & - \\ 107 & 15-30 & 0. & - & 40. & 180 & 2.5 & 3.9 & - \\ 108 & 15-30 & 0.-0.5 & 0-360$^o$ & 10. & - & - & 4.6 & - \\ 109 & 15-30 & 0.-0.5 & 0-360$^o$ & 20. & 50 & 2.8 & 4.5 & - \\ 110 & 15-30 & 0.-0.5 & 0-360$^o$ & 30. & 130 & 2.8 & 4.3 & - \\ 111 & 15-30 & 0.-0.5 & 0-360$^o$ & 40. & 180 & 2.6 & 3.7 & - \\ 112 & 15-30 & 0.-0.5 & 155$^o$ & 10. & - & - & 4.6 & 2.9 \\ 113 & 15-30 & 0.-0.5 & 155$^o$ & 20. & 50 & 2.2-4.0 & 4.5 & 2.8 \\ 114 & 15-30 & 0.-0.5 & 155$^o$ & 30. & 110 & 2.7-3.7 & 4.3 & 3.5 \\ 115 & 15-30 & 0.-0.5 & 155$^o$ & 40. & 180 & 2.7-3.4 & 3.7 & 4.0 \\ \hline \end{tabular} \end{table*} \begin{figure}[tbp] \ifthisfig \psfig{file=u06.ps,height=\columnwidth,rheight=8.5truecm} \else\picplace{6truecm}\ifnofig\thisfigfalse\else\thisfigtrue\fi\fi \caption[]{Plot of the surface brightness $F(r)$ calculated from run \# 106. There is an abrupt change in the slope at $r=100$~AU. The slope is $\alpha=-2.7$ inside and $\alpha=-4.3$ outside this limit. For comparison, the dashed lines represent the power laws $r^{-2.7}$ and $r^{-4.3}$.} \label{106} \end{figure} \subsection{Asymmetry} The distribution observed at large distances in $\beta\:$Pictoris\ disk is obviously not axisymmetric (Smith \& Terrile 1987, Kallas \& Jewitt 1995). However, relative keplerian motions should remove such asymmetries: for two particles on circular orbits with semi-major axis $a_1$ and $a_2$, the time needed to put them in opposite side is $t\sim 0.4(a_1^{-3/2}-a_2^{-3/2})^{-1}$~years if $a_1$ and $a_2$ are in AU. With $a_1=100$~AU and $a_2\ge400$~AU we have $t\le 450 $~years, an extremely short time compared to the age of the system. Thus, any asymmetry should quickly disappear. As noted above, it is however possible to find the asymmetry observed between the two extensions of the $\beta\:$Pictoris\ disk by assuming a peculiar parent body distribution (see Fig.~\ref{113}). Runs \# 112 to \# 115 have been obtained with a peculiar longitude of periastron of the parent bodies ($\omega=155^o$). We found a brightness ratio between the two extensions of the disk ranging from 2.8 to 4.0. This result shows that the observed ratio of SW to NE extension brightness from 1.1 (Lecavelier des Etangs et al. 1993) to 1.5 (Kallas \& Jewitt 1995) can be reproduced if there is a small additional proportion of bodies with peculiar longitudes of periastron like in the FEB scenario. However, there is another way to obtain the observed asymmetry, namely to assume that bodies with different longitudes of periastron move inwards with different velocities. For example, if bodies with a given longitude of periastron move more slowly inward, it produces a fainter disk extension with a steeper brightness gradient in one particular direction (Fig.~\ref{108+115}). \begin{figure}[tbp] \ifthisfig \psfig{file=u13.ps,height=\columnwidth,rheight=8.5truecm} \else\picplace{6truecm}\ifnofig\thisfigfalse\else\thisfigtrue\fi\fi \caption[]{Plot of the surface brightness of the two extension calculated from run \# 113. In this run all the parent bodies have a longitude of periastron of $155^{o}$ relative to the line of sight. The brightness ratio between the two extensions is equal to 2.8. For comparison, the dashed lines represent the power laws $r^{-2.2}$ and $r^{-4.5}$.} \label{113} \end{figure} \begin{figure}[tbp] \ifthisfig \psfig{file=u16.ps,height=\columnwidth,rheight=8.5truecm} \else\picplace{6truecm}\ifnofig\thisfigfalse\else\thisfigtrue\fi\fi \caption[]{Plot of the surface brightness calculated from the addition of runs \# 108 and \#115. In this combined run the parent bodies with a longitude of periastron of $155^{o}$ go more slowly inwards. The brightness ratio between the two extensions is equal to 1.3. The fainter extension has a steeper slope and is represented by a long-dashed line.\\ As in Fig.~1., the hatched zone represents the distribution range of slopes deduced by different observers around $\beta\:$Pictoris.} \label{108+115} \end{figure} \subsection{Mass of the disk} \label{mass} The total mass of the $\beta\:$Pictoris\ disk calculated from optical observations is \begin{center} \begin{equation} M=\int\int n_s(r,s) \frac{4\pi }{3}\rho s^3 2\pi r dr dn_s \end{equation} \end{center} where $s$ is the particle size and $dn_s$ the size distribution. As discussed above we can take $dn_s=s^{-3.5}ds$. Previously, authors assumed that $s$ is independent of $r$ the distance from the star. This assumption is the simpler one and others must be justified. This means however that: \begin{center} \begin{equation} M\propto \frac{8\pi ^{2}\rho }{3} \int_{r_{in}}^{r_{out}}n_s(r)rdr \int s^{-0.5}ds \end{equation} \end{center} The second integral diverges towards the largest bodies; the first one also diverges for $r_{out} \rightarrow \infty$ if the slope of the scattering light is $\ge-4$ (that is the case according to Artymowicz et al. (1989), Lecavelier des Etangs et al. (1993) and for Golimowski et al.'s (1993) NE extension). Thus to evaluate the mass, an arbitrary limit for the disk extension and particle size must be fixed. In the framework of our model both problems are directly solved. At a given distance $r$, we have $ 2\epsilon \le s \le \frac{2\epsilon r}{r-a_0} $ where $a_0$ is the distance of the parent body to the star and $\epsilon$ is defined by $\epsilon=s\beta $. We saw in Section~\ref{size} that $\epsilon$ can be assumed to be constant. We obtain \begin{center} \begin{equation} M \propto \frac{4\pi \rho}{3}\int \left( n_s(r)2\pi r dr \int_{2\epsilon}^{(2\epsilon r)/(r-a_0)}s^{-0.5}ds\right) \end{equation} \end{center} \begin{center} \begin{equation} M \propto \frac{16\pi ^2}{3}\rho \sqrt{2\epsilon} \int_{r_{in}}^{\infty} n_s(r)\frac{a_0}{2}dr \end{equation} \end{center} This integral now converges and the computed mass is of the order of few lunar masses, i.e. of the same order as the lower limit evaluated by Artymowicz et al. (1989) who assumed an arbitrary outer limit of 500~AU and single-size particles of $1\mu$. This mass is consistent with the total mass deduced from submillimeter observations (Zuckerman \& Becklin 1993). \subsection{Gas-dust ratio and CO detection.} \label{CO} One can investigate the connection between the dust observed around $\beta\:$Pictoris\ since 1984 and the gas disk with its stable component thanks to the recent detection of CO in UV lines with the HST ($N_{co}\approx 10^{15}$~cm$^{-2}$, Vidal-Madjar et al. 1994). The very presence of CO may also need a permanent replenishment naturally provided by evaporation of comet-like bodies. We can evaluate the total mass of dust $M_d$ which is associated with the evaporation of CO: \begin{center} \begin{equation} M_d=M_{co} \varphi \psi _{dust/gas} \psi _{gas/CO} \frac{t_d}{t_{CO}} \end{equation} \end{center} where $M_{CO}$ is the total mass of CO in the disk, $\psi_{dust/gas}$ and $ \psi_{gas/CO}$ are the mass ratios of dust to gas and gas to CO, $\varphi$ is the mass ratio of the dust produced effectively kept in the disk and $t_d$ and $t_{CO}$ are the life time of dust and CO: due to photodissociation by UV interstellar radiation $t_{CO}\sim 300$~years (Vidal-Madjar et al. 1994), and at a distance of 100 AU where CO begins to be vaporized, $t_d \sim 10^6$~years (Backman \& Paresce 1993). {}From the upper limit to particle size ($\beta \ge 0.4$), one can evaluate $\varphi\sim 0.1$. With a simple geometry of a disk with an opening angle of 10$^o$, we can deduce from the observed CO column density a total CO mass of about $M_{co} \approx 2\cdot 10^{20}$~kg. {}From observations in the Solar System, we can take $\psi_{dust/gas} \ge 0.1$ (Newburn \& Spinrad 1989). Finally, $\psi_{gas/CO}$ depends on the composition of the evaporated gas; it is between 1. if only CO is present and $\sim 10.$ if all volatiles evaporate (Mumma et al. 1993). Thus, with $\varphi \psi _{dust/gas} \psi _{gas/CO} \approx 0.1$, $M_d$ is about one lunar mass, that is of the same order as the total mass of the dust disk independently evaluated! This rough calculation shows that CO and dust are compatible with a common origin and produced by the same process: evaporation of comet-like bodies at large distances from the star. \section{Conclusion} \label{conclusion} The model we have proposed is able to account for the main characteristics of the $\beta\:$Pictoris\ disk as natural consequences of the production of secondary origin particles by small bodies. It is in fact an answer to the question 3 of Zuckerman \& Becklin (1993): are the dust grains primordial or continually replenished? Indeed, we have shown that if bodies produce dust particles with a given size distribution, these particles follow very eccentric orbits. Thus, a disk can be seen at large distances from the star, and the spatial distribution of dust is very close to a power-law. Moreover, the particles have the same inclinations as the parent bodies: the thickness of the disk thus increases with radius, and the disk is a "wedge" disk: (issue 4). The asymmetric spatial distribution can be explained if we assume that the parent bodies distribution is not axisymmetric (issue 2). This assumption is not surprising since asymmetry is observed for the FEBs and is probably the case in the Solar System for Kuiper belt objects trapped in planetary resonances (Jewitt \& Luu 1995). Since particles are produced with periastrons greater than or equal to those of the parent bodies, the dust present close to the star is either produced there or brought in by Poynting-Robertson drag: the central part must be relatively clear (issue 3). Finally, this model gives connection between the particle size and their distances from the central star, and enable us to solve the issue of the disk mass (issue a), Section~\ref{mass}). Thus, the issues 2), 3), 4) and a) listed in Section~\ref{intro} are simultaneously solved with the simplest form of the model presented here. However, in order to explain the flatter gradient observed in the $\beta\:$Pictoris\ disk than the one obtained without additional hypothesis, we propose to add two other assumptions:\\ 1) The dust is produced by evaporation of bodies of few kilometers in radius, so that there is a balance between evaporating gas drag and gravitation of these parent bodies, and only the smallest particles are extracted.\\ 2) These bodies are moving slowly towards $\beta\:$Pictoris, so that they become extinct before arriving close to the star where the evaporating rate is large enough to produce the largest particles. \\ This OEB (Orbiting-Evaporating-Bodies) scenario is proposed to solve issue 1), but surprisingly it is also able to explain the abrupt change in the slope by the cutoff in distribution of particle size (issue 5)). It has to be noted that the cutoff must not necessary be very sharp and in fact is rather smooth in the model since we assumed a parent body size distribution. The model gives also a natural connection between the dust and gas disks (issue b): they may well be produced by the same bodies in the same process. The connection between the FEB and asymmetry in the disk seems to be possible (issue c)), however this point needs further analysis. This new model is able for the first time to explain simultaneously these issues, but we must now ask if other models or hypothesis can also be made with the same consequences. If correct, this OEB scenario allows to see the $\beta\:$Pictoris\ disk as a gigantic multi-cometary tail with all its components: gas and dust. \begin{acknowledgements} We would like to express our gratitude to Dr. F.Roques who first recall us that ejected particles follow peculiar eccentric orbit. We are particularly indebted to the anonymous referee for his very useful comments which substantially improved the paper. Our thanks go also to Dr. M. Friedjung for improving the manuscript. \end{acknowledgements}
\section{Basic Method} Previous numerical simulations have been performed in a similar way, essentially aimed at producing catalogues of galaxies. Chokshi {\it et al.} \cite*{chokshi88} simulate galaxies in clusters uniformly distributed in space. In each cluster, about 1,000 galaxies are drawn, all at the same redshift. The magnitudes are sorted according to the Schechter Luminosity Function (LF), brighter than a certain cut-off which of course depends on the cluster redshift. Yee \cite*{yee91} first sorts the apparent magnitude according to the observed distribution, the redshift is then computed in bins of apparent magnitude according to a certain Luminosity Function (LF) based on a given, fixed, cosmological and evolutionary model. Our method is slightly different: first, a functional form for the evolution is chosen, and the cosmological parameters are varied on a certain grid. There is then a unique evolution parameter ensuring the condition $<V/V_{a}> = 1/2$ for a given, actual, sample, $V_{a}$ being the available comoving volume according to Avni {\it et al} \cite*{avni81}. The zero--redshift LF $\Phi $ is fixed by such a procedure, and fitted according to some previous analytic choice for the LF. Once the evolution of the LF is known, $V_{a}(M_0)$ may be computed in terms of the zero--redshift absolute magnitude $(M_0)$. The resulting probability distribution function {\it in the sample} of absolute magnitude is then $\Phi (M_0)\times V_{a}(M_0)$. It is easy to sort the absolute magnitude of object $i$ according to this PDF. A second parameter, either the apparent magnitude or the redshift, is needed to define a QSO. Given the absolute magnitude $M_0$ and the cosmological and evolutionary model, $V_a(M_0)$ may be computed. A second random drawing according to a uniform distribution gives a number $x \in [0,1]$, and the redshift of object $i$ may be computed from the condition: $V=x \times V_{a}(M_0) $. $(m,z)$, the apparent magnitude--redshift pair, is then derived from these two variables. For each cosmological model, the procedure is repeated until the desired catalogue size is reached. The null hypotheses are: that PLE applies, that the luminosities are distributed according to the observed LF and that, according to the Cosmological Principle, the QSOs are uniformly distributed in volume on Gpc scales, so that the variable $V/V_{a}$ is uniformly spread over $[0,1]$. From the LF we are able to draw the two variables $M_0$, the zero--redshift absolute magnitude, and $x$; and from the assumption of uniformity we draw the $V/V_{max}$ ratio. Within this framework, particular attention must be paid to the computation of individual $V/V_{a}$. \section{Theoretical background} \subsection{Cosmology } The universe is described by a Friedman-Robertson-Walker model which is defined by three independent fundamental parameters (calculated at $z=0$): the matter density $\Omega_{m}$, which may include a non-baryonic component, the cosmological constant $\Omega_{\Lambda}$, and the Hubble parameter $H_0$. The value of $H_0$ is very important for the creation of synthetic catalogues because it scales the distance and the absolute luminosity of the objects. Nevertheless, $H_0$ has no influence on individual $V/V_{a}'s$, which are scale free. The other known cosmological parameters are related to these three parameters. No arbitrary choice of a preferred cosmological model is made. General formulae for the distance and the volume are given in Caroll {\it et al.} \cite*{caroll92}. We define the conformal coordinate $\chi (z)$ as: \begin{eqnarray} \chi (z)&=&\, 2\, |\Omega_K|^{1/2}\,\, \times \nonumber\cr &&\nonumber\cr && \quad \, \int_0^z {dz'\over ((1+z')^2(1+\Omega_m z')-z'(2+z')\Omega_{\Lambda})^{1/2}} \quad ; \label{integ} \end{eqnarray} \noindent The volume out to redshift $z$ is written: \begin{equation} V(z)=\left({c\over {H_0}}\right)^3{1\over 2\Omega_K|\Omega_K|^{1/2}}\left[ sinn(\chi(z))-\chi(z) \right], \label{volu} \end{equation} \noindent where $\Omega_K=1-\Omega_m-\Omega_{\Lambda}$ is the curvature, $sinn$ is $sin$ if $\Omega_K<0$ and $sinh$ if $\Omega_K>0$. This equation is very useful for numerical computation because it does not require inverse trigonometric functions. Contrary to what happens in a closed universe, with the expressions given, e.g., in Caroll {\it et al.} \cite*{caroll92}, Eq.(\ref{volu}) is single valued and there is no discontinuity at the value $\chi =\pi $, as shown in Mathez {\it et al.} \cite*{mathez91}. \subsection{Evolution} The quasar population resides in an extremely young universe, and it has evolved up until the present epoch. It is likely that individual QSOs do not live for a long period of time, so we speak rather about the evolution in a statistical sense. In any case, assuming a PLE model for quasars amounts to assuming that the fraction of active quasars is constant with epoch, their luminosity decreasing with increasing time, and that they are in sufficient number so that we may apply this average evolutionary law to individual quasars in order to derive maximum redshifts. The absolute magnitude at redshift $z$ is given by: \begin{equation} M(z)=M_0-2.5log_{10}(e(z)), \label{evolu} \end{equation} \noindent where $M_0$ is the absolute magnitude at $z=0$, and $e(z)$ is the evolutionary law: \begin{equation} e(z)\equiv (1+z)^{k_L} \label{pwle} \end{equation} for a Power Law Luminosity Evolution (PWLE), and \begin{equation} e(z) \equiv exp\left(k_L t(z)/H_0\right) = exp\left(t(z)/\tau \right) \label{lexp} \end{equation} for an Exponential Luminosity Evolution (LEXP); $k_L$ is the evolutionary parameter, $t(z)$ the look-back time and $\tau $ the characteristic evolutionary time. Strictly speaking, as quoted by Bigot and Triay \cite*{bigot91}, this evolution law has only a statistical sense and does not apply to individual QSOs. However, to compute each $V/V_{a}$ value, we fictitiously move each QSO toward high redshifts, and it is necessary to apply this evolution law to each QSO. Assuming a power law spectrum $f_\nu=\nu ^{-\alpha}$ with a spectral index $\alpha=0.5$, the Mattig relation relates the observables $(m,z)$ of a quasar to $M_0$, its absolute magnitude at $z=0$: \begin{equation} m_{M_0}(z)=M_0-5+2.5log_{10}\left({d_L^2(z) K(z)\over e(z)}\right); \label{mattig} \end{equation} \noindent here $K(z)= (1+z)^{(\alpha-1)}$ is the K-correction. \begin{figure*} \picplace{20cm} \caption{% Four possible $apparent\; magnitude-redshift$ plots: monotonicaly increasing, one maximum, both one maximum and one minimum, and monotonicaly decreasing. The evolutionary tracks are drawn according to the cosmology and exponential luminosity evolution as described by the top labels. The $m=m(z)$ curves (solid lines) correspond to bins of constant width in absolute magnitude. The horizontal dashed line represents a constant limiting magnitude $B=21$; the dashed $m(z)$ line corresponds to the faint separator ${\cal S}_2$ between magnitude-limited and redshift-limited quasars (see text). Depending on the parameters (cosmology and evolution), the same quasar may be either magnitude-- or redshift--limited. In models where $m(z)$ shows a maximum (for example in Fig. 1b), for quasars lying between $\cal S$ and $B=21$ there exists a redshift slice where $m(z)$ is above $B=21$. The corresponding volume must be subtracted from the available volume $V_2-V_1$. Quasars of the complete Boyle {\it et al.} (1990) sample are superimposed. } \end{figure*} The evolutionary tracks $m(z)$ on the Hubble diagram are shown in Fig. 1 for a given set of quasars in four different cosmologies and evolutionary histories. The curves appear to be quite different. In particular the quasars in Fig1.b,c,d become brighter and brighter as the redshift increases, because the evolution dominates the cosmological dimming at large $z$. \subsection{Distribution of the QSOs in the Hubble plane} Available complete QSO samples usually have a well defined limiting magnitude, towards faint magnitudes. Due to saturation effects (in the photometry or in the spectroscopy), a bright limiting magnitude may also apply. All UVX QSOs are limited to redshifts $z\le 2.2$, precisely because of their selection as objects with a UV excess, a standard way to detect quasars in optical astronomy. As the redshift decreases (below $z=0.3$), there is another limitation because quasars become spatially resolved and can no longer be classified as stellar objects. These redshift limitations prevent the use of maximum (respectively minimum) redshifts outside the limiting redshifts of the sample. Assume that we have a QSO sample in the magnitude and redshift ranges $[m_{1},m_{2}]$ and $[z_1,z_2]$, respectively. For $UVX$ samples we would have $z_1=$0.3 and $z_2=$2.2. We define $\Phi (M,z)$ as the LF at redshift $z$ and at absolute magnitude $M$. However, for the case of PLE it can be written $\Phi (M(z))$; then all luminosity functions at different redshifts differ by an additive constant along the absolute magnitude axis. This allows us to work with the LF defined at a fixed epoch ($z=0$) where $\Phi (M(z=0))=\phi (M_0)$, whatever the evolutionary law \cite{kassiola91}. The number of objects in the range of absolute luminosity $[M_0,M_0+dM_0]$ and redshift $[z,z+dz]$ writes: \begin{eqnarray} d^2N=\phi (M_0) \, dM_0 \, H(m_2-m) \, H(m-m_1)\;\; \times \hfill \nonumber\cr \hfill \, H(z_2-z) \, H(z-z_1) \, {dV\over dz}, \, \, \, dz \;\;\; , \label{dene} \end{eqnarray} \noindent with the following notations: \begin{itemize} \item $dV$ is the comoving volume element of the shell of redshift $z$ to $z+dz$ in the solid angle $\omega $ of the survey; \item $\phi (M_0)$ is defined above; \item The four Heaviside functions are the selection function, ensuring that apparent magnitudes and redshifts remain inside 'box $\cal B$', or $[m_1,m_2]\otimes[z_1,z_2]$, the limitations due to the selection criteria. \end{itemize} We now define ${\cal D}$ as the domain in the Hubble plane where all quasars are redshift--limited. Let $z_{top}$ (respectively $z_{bot}$) be the redshift in the range $[z_1,z_2]$ at which the evolutionary track passes through the maximum (minimum) of the $m(z)$ curve. $z_{top}$ ($z_{bot}$) is the same for all quasars, since their $m(z)$ curves are parallel, differing by an additive constant equal to the absolute magnitude. Consider a given absolute magnitude $M_0$, and let $m_{M_0,top}$ ($m_{M_0,bot}$) be the corresponding faintest (brightest) apparent magnitude in box $\cal B$: \begin{eqnarray} m_{M_0,top}&=\sup_{z\in [z_1,z_2]}\left[ m_{M_0}(z) \right]&= m_{M_0}(z_{top});\nonumber\cr \\ m_{M_0,bot}&=\inf_{z\in [z_1,z_2]}\left[ m_{M_0}(z) \right]&= m_{M_0}(z_{bot}). \label{topbot} \end{eqnarray} $m_{M_0}(z)$ is the apparent magnitude corresponding to the absolute magnitude $M_0$ at redshift $z$. $z_{top}$ is equal to $z_2$ in the case of little or no luminosity evolution. Consider the $m(z)$ curve corresponding to $m_{top}=m_{2}$: it is unique; let us call it the 'faint separator' ${\cal S}_2$. There is also a 'bright separator' -- the unique curve ${\cal S}_1$ crossing $(z_{bot},m_{bot}=m_1)$. ${\cal S}_2$ is drawn on Fig. 1. Domain ${\cal D}$ is simply the domain located {\it below } ${\cal S}_2$ and {\it above} ${\cal S}_1$. Let $M_0^{S_i}$ be the two absolute magnitudes corresponding to the curves ${\cal S}_i, \; (i=1,2)$. The domain ${\cal D}$ simply consists of that part of the Hubble plane brighter than $M_0^{S_2}$ and fainter than $M_0^{S_1}$. We call $\cal D^*$ the complement of $\cal D$ with respect to the box $\cal B$. We call $\cal D$ and $\cal D^*$ quasars the quasars located in each of these domains. \subsection{The $V/V_{max}$ test} Since our method relies on $V/V_{a}$ values, it is necessary to review the $V/V_{a}$ computation, particulary in the case of strong evolution. Consider a quasar in the sample with $(m,z)$ as its representative point in the Hubble diagram, and define the following quantities: \begin{itemize} \item $M_0$: the quasar absolute magnitude at epoch $z=0$; \item $z_{max}(M_0,m_i) \;(i=1,2)$: the redshift(s) where the quasar has an apparent magnitude $m_i$; \item $V_{max}(M_0,m_i)$: the corresponding comoving volume(s); \item $V_{a}$: the available volume; \item $V_i \; (i=1,2)$: the comoving volumes enclosed by the two limiting redshifts $z_i$. \end{itemize} The various possible shapes of the evolutionary track (Fig. 1) are crucial for understanding the $V/V_{a}$ test. The quasar may leave the sample in one of two ways: either its magnitude leaves the range $[m_1,m_2]$, or its redshift leaves the range $[z_1,z_2]$. The available volume $V_{a}$ does {\it not} reduce to $V_{max}(M_0,m_2)$: it is the total volume $V_2-V_1$, minus $\sum V_{out}$, the sum of the volumes of all slices of the universe where the evolutionary track is outside box $\cal B$, so that it prevents the quasar from remaining in the sample. Assuming that volumes are randomly distributed over the range $[V_1,V_2]$, the variable uniformly distributed over $[0,1]$ is the ratio $x$ defined below. The number of quasars in the redshift range $[z_1,z_2]$ is: \begin{eqnarray} N&=&\int_{(M_0)_1}^{(M_0)_2}dM_0 \Phi (M_0) \int_{sup(z_1,z_{max}(M_0,m_1)}^{inf(z_2,z_{max}(M_0,m_2))}dz{dV\over dz} \nonumber\cr \\ &=&\int_{(M_0)_1}^{(M_0)_2}dM_0 \Phi (M_0)V_{a}(M_0), \label{numb} \end{eqnarray} where: \begin{eqnarray} V_{a}(M_0)&=&inf(V_2,V_{max}(M_0,m_2)) \;\; - \cr && sup(V_1,V_{max}(M_0,m_1))\;-\;\sum V_{out} . \label{available} \end{eqnarray} \noindent Note that for $\cal D$ quasars $ \; inf(z_2,z_{max})=z_2$ and $sup(z_1,z_{min})=z_1$ {\it simultaneously}. It is straightforward to find the new variable $x$: \begin{equation} x=(V-V_{sup}-\sum V_{out})/(V_{inf}-V_{sup}-\sum V_{out}). \label{nvar} \end{equation} \noindent Indeed, one shows that the mean of the ratio $x$ over the sample is $1/2$: \begin{eqnarray} <x>&=&{1\over N} \int_{(M_0)_1}^{(M_0)_2}dM_0 \Phi (M_0)\;\;\times \cr && \int_{V_{sup}}^{V_{inf}} dV {(V-V_{sup}-\sum V_{out})\over (V_{inf}-V_{sup}-\sum V_{out})} \nonumber\cr &&\nonumber\cr &&\nonumber\cr &=&1/2, \qquad \forall \Phi , \forall (M_0)_1, \forall (M_0)_2; \label{undemi} \end{eqnarray} \noindent moreover, since $dV=V_{a}(M_0) dx$, we have: \begin{eqnarray} dN&=&\int_{(M_0)_1}^{(M_0)_2} dM_0 \Phi (M_0)V_{a}(M_0) \; dx\nonumber\cr &&\nonumber\cr &&\nonumber\cr &=&N \;\; dx, \label{unif} \end{eqnarray} \noindent which characterizes a PDF of $x$ which is uniform over $[0,1]$. For a redshift--limited sample over $[z_1,z_2]$, taking $V_{sup}=0$ overestimates $V/V_{a}$ and taking $V_{inf}=V_{a}$ underestimates $V/V_{a}$. In either case, $<x>$ cannot be equal to $1/2$ {\it even in the correct cosmology and with the correct evolutionary law.} Varying the shape of the evolutionary track changes the expression of the variable $x$. There may exist zero, one or several maxima of $m(z)$, or the quasar may become fainter than $m_2$ at low redshift; in fact there may be several roots of the equation $m(z)=m_2$. Fig. 1 shows various curves $m(z)$ for various cosmologies and evolutionary histories, in order to illustrate the existence of redshift ranges, lying inside the range $[z_1,z_2]$, but for which $m(z)$ is {\it not} in the range $[m_1,m_2]$. Such ranges must be excluded from the volume computation. \section{ Luminosity Functions in real samples} The LF is not a direct observable -- on the contrary it depends on the choice of the cosmological model and of the evolutionary law. For our purpose, it is necessary to be realistic in reproducing the data; in particular, the artificial catalogues must have the same LF as the real sample whatever the choice of their parent cosmological model. As discussed in Kassiola and Mathez, \cite*{kassiola91}, given the cosmology and the evolution, there are two ways to derive the LF from data: firstly one can construct the {\it Global Luminosity Function} (GLF), by shifting the absolute magnitudes of the whole sample to the present epoch $z=0$, according to the evolution law, and secondly one can compute the {\it Restricted Luminosity Functions} (RLFs), i.e. the LFs in different redshift bins. A necessary condition for PLE to apply is that the LF in all redshift bins differ by a simple shift along the absolute magnitude axis. To do this, one has to compare the RLFs. Assuming that PLE is correct, the reason for choosing the GLF to construct the catalogues is to avoid the noise created by binning the data in redshift. In Section 4.3, we test the PLE hypothesis by comparing of the RLF and the GLF. \subsection{ Real data: the Boyle {\it et al.} sample } The complete sample of $UVX$ quasars of Boyle {\it et al.} (1990) is used. It contains 383 quasars, if we exclude the Narrow Emission Lines (NL) and the redshifts outside the range $[0.3,2.2]$. \subsection{The Global Luminosity Function and its derivation} We now consider the computation of the GLF for any redshift--and magnitude limited--QSO sample. The sample is binned in absolute magnitude $M_0$ with bin width $\Delta M_0$. Let $N_i$ be the QSO number in the $i^{th}$ absolute magnitude bin. The GLF in bin $i$ is given by: \begin{equation} \phi _{obs}^i (M_0)=\sum_{j=1\atop M_{0j}\in B_i}^{N_i} \left(1/ V_a^j\right), \label{fctlum} \end{equation} \begin{eqnarray} B_{i}= \left[M_0-{\Delta M_0\over 2},M_0+{\Delta M_0\over 2}\right], \end{eqnarray} \noindent with an error on $\phi$: \begin{equation} \sigma_\phi^i=\left(\sum_j^{N_i} {V_a^j}^{-2} \right)^{1/2}. \label{erreur} \end{equation} \noindent For the case of the Boyle sample, none of the eight fields intersect; thus, the available volume follows directly from Eq. (8): \begin{eqnarray} V_a&=&\sum_k^{N_{fields}} \omega_k \;\; \times \cr && \left[ inf(V_{max}^k,V_2)-sup(V_{min}^k,V_1)- \sum V_{out} \right], \label{vola} \end{eqnarray} \noindent where $\omega_k$ is the solid angle of the $k^{th}$ field. The GLF looks similar for all of the cosmologies we tried. Special attention has been paid to some of the cosmological models (e.g. $\Omega =0, \Lambda=0;\; \Omega =1, \Lambda=0;\; \Omega =0, \Lambda=0.8$), by various authors which found that the LF is well fitted by either a single \cite{marshall83} or a double \cite{boyle88} power--law. We adopted the double power law--model in all cosmologies and for the two functional forms of evolution: \begin{equation} \phi _{model} (M_0)={\phi^\star\over 10^{0.4(M_0-M_\star)(\alpha+1)}+10^{0.4(M_0-M_\star)(\beta+1)}}; \label{funcform} \end{equation} \noindent $M_\star$ is the characteristic absolute magnitude of the two power--law distribution, i.e. the knee of the distribution. Absolute magnitudes for the quasars are computed according to Eq. (\ref{mattig}). The evolution parameter $k_L$ is computed by setting $<x>=1/2$ for the whole sample. Such a procedure differs from current work \cite{marshall83,boyle88}, in which $k_L$ is fitted like the other parameters. The free parameters $\phi^\star$, $M_\star$, $\alpha$, and $\beta$ are derived by a $\chi^2$ minimization of the binned GLF: \begin{equation} \chi^2=\sum_{i=1}^{N_{bin}} \left({\phi_{obs}^i-\phi_{model}\over \sigma_{\phi_{obs}}^i} \right)^2, \label{chi2} \end{equation} \noindent where $N_{bin}$ is the number of absolute magnitude bins. An important difference with other LF computations is that we have only four free parameters. \begin{figure} \picplace{8cm} \caption{Global Luminosity Functions in four cosmologies, for a power--law luminosity evolution. The models $a,b,c,d$ of Table 1 progress from the larger to the thicker line.} \end{figure} \begin{figure} \picplace{8cm} \caption{Same as above, exponential luminosity evolution. The models $e,f,g,h$ of Table 1 progress from the larger to the thicker line.} \end{figure} Table 1 and Figs. 2 and 3 show a panel of GLFs computed for four different cosmologies, scaled to $H_0=50 \; km s^{-1} Mpc^{-1}$. The general behaviour of the GLF is identical for the exponential and power--law evolutions. The GLF is enlarged and QSOs are brighter for a $\Lambda $--dominated universe. This is due to the fact that, at fixed redshift, the distance in this type of model is greater than in other models, and therefore the absolute luminosities are larger. The effect is in the opposite sense for matter--dominated universes where the objects become fainter and the GLF narrower. The knee of the luminosity distribution evolves according to the preceding remarks. The faintest, shortest and densest GLF is obtained for $\Omega=1$ and $\Lambda=0$. The evolution becomes weaker in $\Lambda$-dominated universes but has a more standard value ($k_L$ around 3--4 in PWLE models) if $\Omega$ is greater than 0.4, regardless of the value of $\Lambda$. Despite our different LF computation method, we find results quite similar to those of Malhotra {\it et al.} \cite*{malhotra95} and Boyle {\it et al.} \cite*{boyle90}. \subsection{ The Restricted Luminosity Functions } Restricted LFs (RLFs) have been computed from the Boyle sample in three different redshift bins ($[0.3,1.0]$, $[1.0,1.7]$, $[1.7,2.2]$). We are interested in checking whether or not the PLE hypothesis is confirmed. If this is the case, any RLF may be deduced from the GLF by a simple shift along the absolute magnitude axis according to the evolution law (Eq. \ref{evolu}). Since we have computed the GLF in the last Section, we are able to compare our RLF to the GLF. This is done by a $\chi^2$--procedure between the GLF and the three RLFs: \begin{equation} \chi^2=\sum_{i=1}^{N_{bin}} \left({\phi_{GLF}-\phi_{RLF}^i\over \sigma_{\phi_{RLF}}^i} \right)^2. \label{chi2} \end{equation} Here, $\phi_{RLF}^i$ and $\sigma_{\phi_{RLF}}^i$ refer to the RLF, computed from the data in the limited redshift bins, and $\phi_{GLF}$ is shifted to this redshift range. The result is shown in Table 1 for four cosmologies and the two evolutionary models. The probabilities $P(>\chi^2)$ are high, and, except for one case (LEXP, $\Omega_{mat}=\Lambda=1$), the test succeeds to better than the 15\% confidence level. This result is in condradiction to those of a similar test done by Kassiola and Mathez (1991), which found strong discrepancies between the GLF and the RLF. However, they interpret these discrepancies as the lack of homogeneity of their composite catalogues. Since our results are consistent with PLE (favoured by Boyle {\it et al.}, 1987), we use this hypothesis to construct our synthetic catalogues. \section{ Evolution characteristic time in various cosmologies } In the exponential model, the characteristic evolution time $\tau $ enters explicitely in Eq. (\ref{lexp}), the expression for $e(z)$. This is not the case for the power law model -- but a characteristic time may be defined as the look--back time to which all luminosities were higher by a factor $e$: \begin{equation} \tau = t\left(z_k = exp(1/k_L) - 1\right) . \label{tau} \end{equation} Table 2 (3) gives the evolutionary parameters $k_L$ for a grid of cosmological parameters in PWLE (LEXP). Table 4 (5) gives the corresponding PWLE (LEXP) characteristic times, computed according to Eqs. (\ref{tau}) and (\ref{lexp}), respectively. These times are in units of the Hubble time $H_0^{-1}$. As is well known, the quasar evolution time is of the order of $H_0^{-1}/(7.7\pm3.5)$ for exponential evolution \cite{mathez76,marshall83,boyle88}, strongly dependent on the cosmological model. For power--law evolution, the characteristic time is $H_0^{-1}/(3.7\pm 0.6)$. Tables 6 and 7 give the same characteristic times as Tables 4 and 5, but in units of the age of the universe in each cosmological model. With the exception of the $\Omega =0, \, \Lambda =1$ model, all ratios in Table 6 are close to 0.30$\pm 0.02$, implying that the characteristic evolution time is about equal to 1/(3.2$\pm 0.2$) of the age of the universe, whatever the cosmological model, in PWLE models. In LEXP models, this ratio is 1/(6.7$\pm $0.9). The characteristic times in Tables 4 and 5, apparently quite dependent of the cosmology, are indeed surprinsingly constant when expressed in terms of the age of the universe. \section{An algorithm for producing quasar catalogues} \medskip \subsection{ The algorithm} Fixing the evolutionary law, the cosmology (including the choice of $H_0$) and the catalogue limits (i.e. $m_1,m_2,z_1,z_2$ which allow the box $\cal B$ and domains $\cal D$ and $\cal D^*$ to be defined), each QSO $(m,z)$ is fully determined by the other pair of variables $(V/V_{a},M_0)$. This is easy to understand from Fig. 1, where the Mattig functions appear. All of our previous discussion shows that under the PLE hypothesis (which was checked on the data), any sample is fully determined from its limiting magnitudes and redshifts, the GLF and a uniform distribution of the $V/V_{a}$. From the probability distribution function (PDF) of absolute magnitudes $M_0$, which is $\phi (M_0) \times V_{a}(M_0)$, and from the $V/V_{a}(M_0)$ ratios uniformly spread over $[0,1]$, it is possible to draw Monte--Carlo catalogues. The algorithm is shown in Fig. 4. Choosing the cosmology and a functional form for the evolution, the first part of the work is to compute the evolutionary parameter from the constraint $<V/V_{a}>=0.5$, using the Boyle sample. At the same time, $V_{a}(M_0)$ is computed and used to obtain the GLF. In the second part, two series of random numbers, $\epsilon_1$ and $\epsilon_2$, are used to extract the $V/V_{a}$ and the absolute magnitude $M_0$ of each object in the synthetic catalogue from the relations: \begin{eqnarray} \epsilon_1&=&V/V_{a}={V-V_{sup}-\sum V_{out}\over V_{inf}-V_{sup}-\sum V_{out}},\nonumber\cr \epsilon_2&=&{\int_{-\infty}^{M_0} dM \phi (M) V_{a} (M) \over \int_{-\infty}^{+\infty}dM \phi (M) V_{a} (M) }. \label{catapar} \end{eqnarray} Obtaining the redshifts and apparent magnitudes is straightforward from a simple inversion procedure. Synthetic catalogues have been produced in the past, where both a LF {\it plus} a redshift histogram were jointly fitted. As already explained, this method suffers from at least two problems: Essentially, the two distributions may be incompatible in different cosmologies. Our method has the advantage of avoiding these problems since we use only one distribution, the GLF, and the other distributions (redshift, magnitude) are derived only after the cosmology and the evolution have been fixed. The algorithm shown in Fig. 4 ensures the coherence of these distributions, moreover it takes the evolution into account very accurately. This is not obviously true in previous methods where it is not clear how the evolution intervenes. In fact, we used the Cosmological Principle to replace the redshift distribution of the standard method by demanding the uniformity of $V/V_{a}$. Moreover, the limiting magnitude and the redshift range for the synthetic sample may be different from their values in the input sample. \begin{figure} \picplace{10cm} \caption{Algorithm for the construction of synthetic catalogues} \end{figure} \subsection{ Redshift and Magnitude Histogrames} The input \cite{boyle90} and output (synthetic catalogue with the same observational biases) redshift histogrames are compared in Fig. 5 for models $b$ and $f$ of Table 1. Similarly, the histogrames of absolute magnitude are compared in Fig. 6. The fit is quite satisfactory for both models; this is true for all models we tried. It is possible to introduce various biases in the synthetic catalogues, depending on which sort of catalogue we want to simulate. For example, for UVX quasars it is well known that the color selection criteria introduce a deficiency of QSOs in the redshift range $[1.5,1.8]$ \cite{boyle88}. Moreover, QSOs are variable objects, which can be simulated by randomly selecting a magnitude variation in a pre--defined distribution function. Noise in the magnitude and redshift measurements may be simulated in the same way. \begin{figure} \picplace{8cm} \caption{Redshift Histogrames: (1): Boyle sample; (2): simulated catalogue, translated 15 upwards, model $b$; (3): idem, translated 35 upwards, model $f$ } \end{figure} \begin{figure} \picplace{8cm} \caption{Absolute Magnitude Histogrames: dashed curve: Boyle sample; full line: simulated catalogue, model $b$} \end{figure} \begin{figure} \picplace{8cm} \caption{Absolute Magnitude Histogrames: dashed curve: Boyle sample; full line: simulated catalogue, model $f$} \end{figure} \section{Conclusion} We have reviewed in detail the $V/V_{max}$ computation in various cosmologies. The Mattig relation is used to understand how to compute the available volume $V_a$ of a QSO, even in complicated cases. Both the Pure Luminosity Evolution (PLE) assumption and the Cosmological Principle allow the variables $(M_0,V/V_{a})$ to be used instead of $(m,z)$ to construct synthetic QSO samples in a coherent way. Some of the previous catalogue constructions were shown to suffer problems. The Global Luminosity Function used for this purpose is derived in a slightly different way from previous calculations since we fixed the evolution parameter by the condition $<V/V_{a}>=1/2$. This significantly changes the value of this parameter, and hence of the GLF. A surprising result is that the QSO characteristic evolution times are constant when expressed in terms of the age of the universe, regardless of the cosmology. The PLE hypothesis was checked on the data, which comfirms previous results \cite{boyle88}. The advantage of such catalogues is that we tightly control all aspects of the sample: the parent cosmology, the strength and the functional form of the evolution, the luminosity function, the magnitude and redshift depths, and the effects of magnitude and redshift biases, all of which influence the results of the new cosmological test introduced in Paper I. In Paper I we apply the test to catalogues with different redshift limitations, which leads us to propose a new observational strategy for the construction of future QSO samples. It is possible to test a catalogue with a wrong hypothesis, for example on the evolution, by assuming a power--law evolution and running the cosmological test with the exponential hypothesis. Moreover, our method is tractable with a Pure Density Evolution hypothesis, provided that density--weighted volume elements $\rho dV$ are substituted for all volume elements $dV$. \begin{table*} \caption[] The LF (Eq. (\ref{fctlum})) in various redshift ranges for four arbitrary cosmologies and two evolution laws. Since we are interested in input parameters to construct catalogues, confidence intervals are not needed. Only the best fit parameters are given. We choose $H_0=50 km s^{-1} Mpc^{-1}$ } \begin{flushleft} \begin{tabular}{ccccccccccccccc} \noalign{\smallskip} \hline \noalign{\smallskip} &&&&&&&&&&&\cr &$\Omega $&$\Lambda $&\hfil $k_L$&\hfil $\alpha$\hfil &\hfil $\beta$\hfi &\hfil $M_{\star}$& $\phi^{\star}$&\hfil $P(>\chi^2)$\hfi &\hfil $P(>\chi^2)$\hfil &\hfil $P(>\chi^2)$\hfil &\hfil $P(>\chi^2) \hfil \cr & &&&&&&$10^{-6} Mpc^{-3}$&$z_1=$0.3 & 0.3&1.0&1.7 \cr & &&&&&&&$z_2=$2.2 & 1.0&1.7 &2.2\cr \hline\noalign{\smallskip} PWLE&&&&&&&&&&\cr a&0&0&2.94&-3.32&-1.38&-23.41&1.8$10^{-6}$&\hfil 0.97\hfil &\hfil 0.20\hfil &\hfil 0.78\hfil &\hfil 0.80\hfil \cr b&1&0&3.60&-3.89&-1.30&-22.10&1.1$10^{-5}$&\hfil 0.99\hfil &\hfil 0.99\hfil &\hfil 0.69\hfil &\hfil 0.92\hfil \cr c&0&1&1.91&-2.43&-1.23&-23.76&2.6$10^{-6}$&\hfil 0.89\hfil &\hfil 0.15\hfil &\hfil 0.88\hfil &\hfil 0.58\hfil \cr d&1&1&3.81&-4.22&-1.37&-22.25&3.2$10^{-6}$&\hfil 0.99\hfil &\hfil 0.92\hfil &\hfil 0.66\hfil &\hfil 0.81\hfil \cr \noalign{\smallskip} \hline \noalign{\smallskip} LEXP&&&&&&&&&&\cr e&0&0&6.65&-3.51&-1.49&-22.18&1.5$10^{-6}$&\hfil 0.99\hfil &\hfil 0.58\hfil &\hfil 0.78\hfil &\hfil 0.40\hfil \cr f&1&0&11.8&-4.61&-1.55&-19.84&6.3$10^{-6}$&\hfil 0.97\hfil &\hfil 0.88\hfil &\hfil 0.57\hfil &\hfil 0.24\hfil \cr g&0&1&1.90&-2.69&-1.49&-24.66&1.2$10^{-6}$&\hfil 0.76\hfil &\hfil 0.16\hfil &\hfil 0.23\hfil &\hfil 0.58\hfil \cr h&1&1&9.71&-4.62&-1.60&-20.13&2.1$10^{-6}$&\hfil 0.99\hfil &\hfil 0.64\hfil &\hfil 0.57\hfil &\hfil 0.01\hfil \cr \noalign{\smallskip} \hline \end{tabular} \end{flushleft} \end{table*} \begin{table} \caption[] Power Law Luminosity Evolution (PWLE) parameter in different cosmologies. The model $\Omega =0, \, \Lambda =1$, which is marked with an asterisk, is close to a universe without a Big--Bang, and not very reliable.% } \begin{flushleft} \begin{tabular}{lllllllllllllll} \noalign{\smallskip} \hline \noalign{\smallskip} $\Lambda $ &$\Omega $& 0.0 & 0.2 & 0.4 & 0.6 & 0.8 & 1.0 \cr 0.0 && 2.94 & 3.16 & 3.28 & 3.44 & 3.50 & 3.60 \cr 0.2 && 2.88 & 3.13 & 3.28 & 3.47 & 3.50 & 3.63 \cr 0.4 && 2.78 & 3.10 & 3.28 & 3.47 & 3.53 & 3.66 \cr 0.6 && 2.66 & 3.06 & 3.28 & 3.47 & 3.60 & 3.69 \cr 0.8 && 2.47 & 3.00 & 3.31 & 3.47 & 3.63 & 3.75 \cr 1.0 && 1.91(*) & 2.94 & 3.35 & 3.50 & 3.66 & 3.81 \cr \noalign{\smallskip} \hline \end{tabular} \end{flushleft} \end{table} \begin{table} \caption[] Same as Table 2 for LEXP, Exponential Luminosity Evolution.} \begin{flushleft} \begin{tabular}{lllllllllllllll} \noalign{\smallskip} \hline \noalign{\smallskip} $\Lambda $ &$\Omega $& 0.0 & 0.2 & 0.4 & 0.6 & 0.8 & 1.0 \cr 0.0 && 6.65 & 8.02 & 9.15 & 10.10 & 11.02 & 11.77 \cr 0.2 && 6.02 & 7.40 & 8.65 & 9.71 & 10.59 & 11.40 \cr 0.4 && 5.28 & 6.84 & 8.15 & 9.28 & 10.15 & 11.03 \cr 0.6 && 4.40 & 6.15 & 7.53 & 8.78 & 9.71 & 10.65 \cr 0.8 && 3.40 & 5.40 & 6.96 & 8.21 & 9.28 & 10.15 \cr 1.0 && 1.90(*) & 4.59 & 6.28 & 7.65 & 8.78 & 9.71 \cr \noalign{\smallskip} \hline \end{tabular} \end{flushleft} \end{table} \begin{table} \caption[] Evolution characteristic times: look--back time to which all luminosities were higher by a factor $e$, given the PWLE parameter in Table 2. Times are in units of $H_0^{-1}$, the mean is 0.27 and the dispersion is 0.04. } \begin{flushleft} \begin{tabular}{lllllllllllllll} \noalign{\smallskip} \hline \noalign{\smallskip} $\Lambda $ &$\Omega $& 0.0 & 0.2 & 0.4 & 0.6 & 0.8 & 1.0 \cr 0.0 && .29 & .27 & .26 & .24 & .23 & .23 \cr 0.2 && .30 & .28 & .26 & .24 & .24 & .23 \cr 0.4 && .32 & .29 & .27 & .25 & .24 & .23 \cr 0.6 && .35 & .30 & .28 & .25 & .24 & .24 \cr 0.8 && .39 & .31 & .28 & .26 & .25 & .23 \cr 1.0 && .52(*) & .33 & .29 & .27 & .25 & .24 \cr \noalign{\smallskip} \hline \end{tabular} \end{flushleft} \end{table} \begin{table} \caption[] Same as Table 4, but for LEXP. Mean characteristic time: 0.13; dispersion: 0.05. } \begin{flushleft} \begin{tabular}{lllllllllllllll} \noalign{\smallskip} \hline \noalign{\smallskip} $\Lambda $ &$\Omega $& 0.0 & 0.2 & 0.4 & 0.6 & 0.8 & 1.0 \cr 0.0 && .15 & .12 & .11 & .10 & .09 & .08 \cr 0.2 && .17 & .13 & .12 & .10 & .09 & .09 \cr 0.4 && .19 & .15 & .12 & .11 & .10 & .09 \cr 0.6 && .23 & .16 & .13 & .11 & .10 & .09 \cr 0.8 && .29 & .18 & .14 & .12 & .11 & .10 \cr 1.0 && .53(*) & .22 & .16 & .13 & .11 & .10 \cr \noalign{\smallskip} \hline \end{tabular} \end{flushleft} \end{table} \begin{table} \caption[] PWLE characteristic times (Table 4) in units of the age of the universe in the corresponding cosmology. Mean: 0.31; dispersion: 0.02. These ratios are far more independent of cosmology than the times in Table 4. } \begin{flushleft} \begin{tabular}{lllllllllllllll} \noalign{\smallskip} \hline \noalign{\smallskip} $\Lambda $ &$\Omega $& 0.0 & 0.2 & 0.4 & 0.6 & 0.8 & 1.0 \cr 0.0 && .29 & .32 & .33 & .33 & .34 & .34 \cr 0.2 && .28 & .31 & .33 & .32 & .33 & .34 \cr 0.4 && .28 & .31 & .32 & .32 & .33 & .33 \cr 0.6 && .27 & .30 & .31 & .31 & .32 & .32 \cr 0.8 && .25 & .29 & .30 & .30 & .31 & .31 \cr 1.0 && .13(*)& .28 & .29 & .30 & .30 & .31 \cr \noalign{\smallskip} \hline \end{tabular} \end{flushleft} \end{table} \begin{table} \caption[] Same as Table 6, but for LEXP. Mean: 0.15; dispersion: 0.02. As for PWLE, the characteristic time depends far less on the age of the universe than on the Hubble time. } \begin{flushleft} \begin{tabular}{lllllllllllllll} \noalign{\smallskip} \hline \noalign{\smallskip} $\Lambda $ &$\Omega $& 0.0 & 0.2 & 0.4 & 0.6 & 0.8 & 1.0 \cr 0.0 && .15 & .15 & .14 & .14 & .13 & .13 \cr 0.2 && .16 & .15 & .14 & .14 & .13 & .13 \cr 0.4 && .16 & .16 & .15 & .14 & .13 & .13 \cr 0.6 && .17 & .16 & .15 & .14 & .13 & .13 \cr 0.8 && .19 & .17 & .15 & .14 & .13 & .13 \cr 1.0 && .13(*)& .18 & .16 & .14 & .14 & .13 \cr \noalign{\smallskip} \hline \end{tabular} \end{flushleft} \end{table} \paragraph{Acknowledgements} We thank B. Fort, P.Y. Longarreti, G. Soucail, J.P. Picat, R. Pell\'o and J.F. Leborgne for support and discussions, and especially Jim Bartlett for a careful reading of the manuscript and enlighting suggestions. L.V.W. thanks the french MESR for grant 93135. This work was supported by grants from the french CNRS (GdR Cosmologie) and from the European Community (Human Capital and Mobility ERBCHRXCT920001).
\section{Introduction} The subject of applications of classical Lie symmetries to boundary value problems is well studied (see the monography \cite{ovs}). In contrast, the question of involving higher symmetries to the same problem has received much less attention, unlike say the Cauchy problem. However, one should stress that nowadays the higher symmetries' approach becomes the basis of the modern integrability theory \cite{mss}. A number of attempts to apply the inverse scattering method (ISM) to the initial boundary value problem has been undertaken. It turned out that if both initial data and boundary value are chosen arbitrary then the ISM loses essentially its efficiency. On the other hand side the investigation by E.Sklyanin \cite{skl} based on the $R-$ matrix approach demonstrated that there is a kind of boundary conditions, compatible completely with the integrability. The analytical aspects of such kind problems were studied in \cite{tar}, \cite{bikt}. After \cite{hab1} it becomes clear that boundary value problems found can effectively be investigated with the help of the B\"acklund transformation. Below we will discuss a higher symmetry test, proposed in \cite{hab2}, \cite{gur} to verify whether the boundary condition given is compatible with the integrability property of the equation. It is worthwhile to note that all known classes of boundary conditions, compatible with integrability occur to pass this symmetry test. Boundary conditions involving explicit time dependance for the Toda lattice compatible with higher symmetries has recently been studied in \cite{adl}. It was established there that finite dimensional systems obtained from the Toda lattice by imposing at both ends boundary conditions consistent with symmetries were nothing else but Painlev\'e type equations. Let us consider the evolution type equation \begin{equation} u_{t}=f(u,u_{1},u_{2},...,u_{n}) \label{eq} \end{equation} and a boundary condition of the form \begin{equation} p(u,u_{1},u_{2},...,u_{k}) \vert_{x=0}=0, \label{gu} \end{equation} imposed at the point $x=0$. Here $u_i$ stands for the partial derivative of the order $i$ with respect to the variable $x.$ Suppose that the equation given possesses a higher symmetry \begin{equation} u_{\tau}=g(u,u_{1},...,u_{m}). \label{sy} \end{equation} We call the problem (\ref{eq})-(\ref{gu}) compatible with the symmetry (\ref{sy}) if for any initial data prescribed at the point $t=0$ a common solution to the equations (\ref{eq}), (\ref{sy}) exists satisfying the boundary condition (\ref{gu}). Let us explain more exactly what we mean. Evidently one can differentiate the constraint (\ref{gu}) only with respect to the variables $t$ and $\tau,$ (but not respect to $x$). For instance, it follows from (\ref{gu}) that \begin{equation} \sum_{i=0}^{n} {\partial p \over \partial u_i}(u_i)_{\tau}=0, \label{dgu} \end{equation} where one should replace $\tau$-derivatives by means of the equation (\ref{sy}). The boundary value problem (\ref{eq})-(\ref{gu}) be compatible with the symmetry (\ref{sy}) if the equation (\ref{dgu}) holds identically by means of the condition (\ref{gu}) and its consequences obtained by differentiation with respect to $t.$ To formulate an effective criterion of compatibility of the boundary value problem with a symmetry it's necessary to introduce some new set of dynamical variables consisting of the vector $v=(u,u_{1},u_{2}, ... u_{n-1})$ and its $t$-derivatives: $v_{t}$, $v_{tt}$, ... . Passing to this set of variables allows one really to exclude the dependance on the variable $x.$ In terms of these variables the symmetry (\ref{sy}) and the constraint (\ref{gu}) take the form \begin{equation} v_{\tau}=G\,(v,v_{t} , v_{tt} ,... {\partial^{m_1}v\over \partial t^{m_1}}), \label{sy1} \end{equation} \begin{equation} P(v, {\partial v\over \partial t},...,{\partial^{k_1}v\over \partial t^{k_1}})=0. \label{gu1} \end{equation} The following criterion of compatibility was established in (\cite{gur}). \th{Theorem.} The boundary value problem (\ref{eq})-(\ref{gu}) is compatible with the symmetry (\ref{sy}) if and only if the differential connection (\ref{gu1}) is consistent with the system (\ref{sy1}). We call the boundary condition (\ref{gu}) compatible with the integrability property of the equation (\ref{eq}), if the problem (\ref{eq})-(\ref{gu}) is compatible with infinite series of linearly independent higher order symmetries. The problem of the classification of integrable boundary conditions is solved completely for the Burgers equation (see \cite{gur}) \begin{equation} u_{t}=u_{2}+2\,u\,u_{1}, \label{bur} \end{equation} \th{Theorem.} If the boundary condition $p(u,u_{1})\vert_{x=0}=0$ is compatible at least with one higher symmetry of the Burgers equation (\ref{bur}) then it is compatible with all even order homogeneous symmetries and is of the form $c_1(u_{1}+u^{2})+c_2\,u+c_3=0.$ In the Burgers case the boundary conditions of the general form (\ref{gu}) can also be described completely with the help of the "recursion operator for the boundary conditions" $L={\partial \over\partial x}+u,$ which acts on the set of integrable boundary conditions (see \cite{svi}). For instance, the boundary condition $L(c_1(u_{1}+u^{2})+c_2\,u+c_3)=c_1(u_{2}+3uu_{1}+u^{3})+c_2(u_1+u^2) +c_3\,u=0$ is also integrable. Let us describe boundary value problems of the form \begin{equation} a(u,u_x)\vert_{x=0}=0 \label{sggu} \end{equation} \begin{equation} u_{tt}-u_{xx}+\sin u=0, \label{sg} \end{equation} for the sine-Gordon equation compatible with the third order symmetry. As it is shown in \cite{jsh} the complete algebra of higher symmetries for the equation (\ref{sg}) i.e. $u_{\xi\eta}=\sin u,$ where $2\xi=x+t,$ $2\eta=x-t$ splits into the direct sum of two algebras consisting of symmetries of equations $ u_{\tau}=u_{\xi\xi\xi}+u_{\xi}^3/2,\, u_{\tau}=u_{\eta\eta\eta}+u_{\eta}^3/2, $ correspondingly, which are nothing else but potentiated MKdV equation. Particularly, the following flow commutes with the sine-Gordon equation \begin{equation} u_{\tau}=c_1(u_{\xi\xi\xi}+u_{\xi}^3/2)+c_2(u_{\eta\eta\eta}+ u_{\eta}^3/2). \label{lcs} \end{equation} The symmetry (\ref{lcs}) isn't compatible with any boundary condition of the form (\ref{sggu}) unless $c_1=-c_2,$ under this constraint the equation (\ref{sggu}) is of one of the forms \begin{equation} u=const, \qquad v=c_1\cos (u/2)+c_2\sin (u/2). \label{sgco} \end{equation} Note that the list of boundary conditions (\ref{sgco}) coincides with that found by A.Zamolodchikov within the framework of the R matrix approach \cite{zam}. The latter in (\ref{sgco}) in particular cases was studied earlier in \cite{skl} and \cite{bikt}. The compatibility of the former in (\ref{sgco}) with the usual version of ISM was declared earlier in \cite{bikt}. But the statement was based in a mistake (see \cite{hab3}). Our requirement of consistency is weaker than that is used in \cite{bikt}. Applications of these and similar problems for the sine-Gordon equation and the affine Toda lattice in the quantum field theory are studied in \cite{sal} and \cite{cor}. According to the theorem above one reduces the problem of finding integrable boundary conditions to the problem of looking for differential connections admissible by the following system of equations, equivalent to (\ref{lcs}) with $c_1=-c_2$ and $v=u_x:$ \begin{equation} \begin{array}{ll} u_{\tau}=8u_{ttt}+6u_{t}\cos u+3v^2u_t+u_{t}^3,\\ v_{\tau}=8v_{ttt}+6v_{t}\cos u+6u_{tt}vu_t+3v^2v_t+3u_{t}^2v_t.\label{sgt2} \end{array} \end{equation} One can prove that the boundary conditions (\ref{sgco}) are compatible with rather large subclass of the sine-Gordon equation such that \begin{equation} u_{\tau}=\phi (u,u_1,...u_{k_1})-\phi (u,\bar u_1,...\bar u_{k_1}), \label{sgsym} \end{equation} where $u_j=\partial^ju/\partial\xi^j,$ $\bar u_j=\partial^ju/\partial\eta^j,$ and the equation $u_{\tau}=\phi_i(u,u_1,...u_{k_i}),$ $i=1,2$ is a symmetry of the equation $u_{\tau}=u_{\xi\xi\xi}+u_{\xi}^3/2.$ Another well-known integrable equation of hyperbolic type \begin{equation} u_{tt}-u_{xx}=\exp (u) +\exp (-2u) \label{ts} \end{equation} has applications in geometry of surfaces. For the first time it was found by Tzitzeica \cite{tzi}. The presence of higher symmetries for this equation has been established by A.Jiber and A.Shabat \cite{jsh}. The simplest higher symmetry of this equation is of the fifth order \begin{equation} u_{\tau}=u_{\xi\xi\xi\xi\xi}+5(u_{\xi\xi}u_{\xi\xi\xi}-u_{\xi}^2u_{\xi\xi\xi} -u_{\xi}u_{\xi\xi}^2)+u^5_{\xi}. \label{tssym} \end{equation} It is proved in the article cited that the symmetry algebra for (\ref{ts}) is the direct sum of the symmetry algebras of (\ref{tssym}) and of the equation obtained from (\ref{tssym}) by replacing $\xi$ by $\eta.$ Let us look for boundary conditions of the form \begin{equation} a(u,u_x)=0, \label{tsgu} \end{equation} for the equation (\ref{ts}), compatible with the symmetry \begin{equation} \begin{array}{lr} u_{\tau}=u_{\xi\xi\xi\xi\xi}+5(u_{\xi\xi}u_{\xi\xi\xi}-u_{\xi}^2u_{\xi\xi\xi} -u_{\xi}u_{\xi\xi}^2)+u^5_{\xi}-u_{\eta\eta\eta\eta\eta}- \\ 5(u_{\eta\eta}u_{\eta\eta\eta}-u_{\eta}^2u_{\eta\eta\eta} -u_{\eta}u_{\eta\eta}^2)-u^5_{\eta}. \label{tst1} \end{array} \end{equation} Rather simple but tediously long computations lead to the following statement. \th{Theorem.} Boundary conditions (\ref{tsgu}) for the Jiber-Shabat equation compatible with the symmetry (\ref{tst1}) (and then compatible with integrability) are either of the form $u_x+c\exp (-u)\vert_{x=0}=0$ or $u_x+c\exp (u/2)\pm \exp (-u) \vert_{x=0}=0 ,$ where $c$ is arbitrary. Notice that all equations above are invariant under the reflection type symmetry $x\rightarrow-x.$ It is unexpected that equations which don't admit any reflection symmetry admit nevertheless boundary conditions compatible with integrability. For instance, the famous KdV equation \begin{equation} u_t=u_{xxx}+6u_xu \end{equation} is consistent with the boundary condition \begin{equation} u=0\vert_{x=0},\, u_{xx}\vert_{x=0}=0. \label{kdvcon} \end{equation} It implies immediately that the boundary value problem $$ u_t=u_{xxx}+6u_xu,\qquad u=0\vert_{x=0} $$ with the Dirichlet type condition at the axis $x=0$ admits an infinite dimensional set of "explicit" finite-gap solutions. This work was partially supported by Russian Foundation of Fundamental Researches (grant 93-011-165) and International Scientific Foundation (grant RK-2000).
\section{Introduction} \label{sec:intro} The formation of a black hole is, in principle, one of the most efficient mechanisms for generation of gravitational waves. Such sources tie together two major research initiatives. Laser interferometric gravity wave detectors\cite{ligo} hold out a promise of the detection of gravitational waves from astrophysical events. To interpret the results of the gravitational wave signals, and to help find signals in the detector noise, a broad and detailed knowledge will be needed of astrophysical gravitational waveforms. This is one of the underlying motivations for the ``grand challenge''\cite{grandchallenge} in high performance computing, aimed at computing the coalescence of black hole binaries. Evolving numerical spacetimes and extracting outgoing radiation waveforms is indeed a challenge. In a straightforward numerical approach, a good estimate of the asymptotic waveform requires long numerical evolutions so that the emitted waves can be propagated far from the source. The necessary long evolutions are difficult for a number of reasons. General difficulties include throat stretching when black holes form, numerical instabilities associated with curvilinear coordinate systems, and the effects of outer boundary conditions which are approximate.\cite{ast95} We suggest here that at least part of the cure for this problem may lie in the use of the theory and techniques of nonspherical perturbations of the Schwarzschild spacetime (``NPS''). By this we mean the techniques for treating spacetimes as deviations, first order in some smallness parameter, from the Schwarzschild spacetime. These techniques differ from ``linearized theory'' which treats perturbations of the spacetime from Minkowski spacetime and which cannot describe black holes. The basic ideas and methods were set down by many authors and lead to ``wave equations'' for the even parity\cite{zerilli} and odd parity\cite{rw} perturbations. NPS has been used to compute outgoing radiation waveforms from a wide variety of black hole processes, including the scattering of waves\cite{waves}, particles falling into a hole\cite{particle}, and stellar collapse to form a hole\cite{stars}. The general scheme of NPS also underlies the techniques for extraction of radiation from numerically evolved spacetimes\cite{ae90}. NPS computations have recently been used in conjunction with fully numerical evolution, as a code test\cite{abhss} and as a strong-field radiation extraction procedure\cite{ast95}. Here we are interested in another sort of application of NPS theory. To understand such applications we consider an example: Two very relativistic neutron stars falling into each other, coalescing and forming a horizon, as depicted in Fig.~1. The curve ``hypersurface," in Fig.~1, indicates a spacelike ``initial" surface. The spacetime can be divided into three regions by this initial surface and the horizon. The early evolution, in region I, below the initial hypersurface, is highly dynamical and nonspherical. Spherical perturbation theory is clearly inapplicable. Above the initial surface the spacetime remains highly nonspherical in region II inside the event horizon, but outside the event horizon, in region III, it may be justified to consider the spacetime to be a perturbation of a Schwarzschild spacetime. This is essentially guaranteed if the initial hypersurface is chosen late enough, in some sense, after the formation of the horizon. The evolution in region III, then, is determined by cauchy data on the initial hypersurface exterior to the horizon. It is important to note that this is made possible by the fact that the horizon is a causal boundary which shields the outer region from the dynamics of the highly nonspherical central region. The scheme inherent in this division of spacetime has the potential greatly to increase the efficiency of the computation of the radiation generated when strong field sources form black holes. If one starts from the cauchy data on the initial hypersurface, one can evolve forward in time with the linear equations of perturbation theory. Many of the long-time evolution problems of numerical relativity are avoided and the interpretation of the computed fields in terms of radiation is immediate. The approach suggested would then seem to be: Use numerical relativity up to the initial hypersurface; use the techniques of nonspherical perturbations in the future of the initial hypersurface. In fact, the efficiency that can be achieved may be even greater. In the early, highly nonspherical, pre-initial hypersurface phase of the development of the spacetime, there may be relatively little generation of gravitational radiation. By using a computational technique which suppresses the radiative degrees of freedom one may be able to compute the early stages of evolution relatively easily. There are two very recent examples of just such applications of this viewpoint. Price and Pullin\cite{price_pullin94} used as initial data the Misner's\cite{misner} solution to the initial value equations for two momentarily stationary black holes. Abrahams and Cook\cite{abrahams_cook94} considered two holes moving towards each other, and used numerical values of the initial value equations. In neither case was there {\em any} use of fully nonlinear numerical evolution. The rather remarkable success of both computations suggests that there is something robust about the underlying idea of separating horizon-forming astrophysical scenarios into an early phase with no radiation and a late phase with small deviations from sphericity outside the horizon. It is plausible that the bulk of the radiation in most processes is generated only in the very strong-field interactions around the time of horizon formation and that radiation generation in the early dynamics can be ignored. One would, however, think that strong radiation would be emitted during the stages at which the early horizon is very nonspherical and at which time nonspherical perturbation theory would seem to be inapplicable. There should be a tendency for this ``early'' radiation, produced very close to the horizon, to go inward into the developing black hole, so that the application of nonspherical perturbation theory to the exterior really requires that on the initial spacetime the perturbation are small only well outside the horizon. It would seem that something of this sort would have to be happening to explain the accuracy of the Price-Pullin and Abrahams-Cook results. Whether or not many problems can be treated with {\em no} use of fully numerical evolution, it appears clear to us that these perturbation methods will be applied to a variety of problems in which data on the initial hypersurface is available numerically. The primary purpose of this paper is to provide justification and background for earlier work on this subject and a clear recipe for future applications. In the next section we discuss the meaning, and limitations, of extracting a ``perturbation'' from this numerical data and computing radiated energies. The explicit process of extracting the perturbations from the numerical data is given in Sec.~III. In Sec.~IV we demonstrate the use of this procedure via application to a specific example, the Misner initial data. \section{Initial data as Schwarzschild perturbations} We outline here the formalism for perturbation theory based on work by Regge and Wheeler\cite{rw} and by Zerilli\cite{zerilli}, but we will draw heavily on the gauge invariant reformulation of those earlier works by Moncrief\cite{moncrief74}. Our starting point is an initial hypersurface which can be taken as a surface of constant Schwarzschild time. We assume that the coordinates ${x^i}$ on that surface are almost Schwarzschild coordinates $r,\theta,\phi$ and we assume that the values are known, on this hypersurface and in these coordinates, for the 3-metric $\gamma_{ij}$ and the extrinsic curvature $K_{ij}$. The conditions for finding such a hypersurface and such coordinates will be made explicit in Sec.~III. Underlying perturbation theory is the idea of a family of metric functions $g_{\mu\nu}(x^\alpha;\epsilon)$, depending on the parameter $\epsilon$, which satisfy the Einstein equations for all $\epsilon$, and which, in the limit $\epsilon\rightarrow0$, become the Schwarzschild metric functions, such as $g_{rr}=S^{-1}$. (Here $S\equiv 1-2M/r$ and $M$ is the mass of the Schwarzschild spacetime; we use units throughout in which $c=G=1$.) NPS theory amounts to the approximation \begin{equation}\label{pertbasic} g_{\mu\nu}(x^\alpha;\epsilon)\approx g_{\mu\nu}(x^\alpha;\epsilon)|_{\epsilon=0}+ \epsilon\frac{\partial}{\partial\epsilon} g_{\mu\nu}(x^\alpha;\epsilon)|_{\epsilon=0}\ . \end{equation} \subsection{Choice of expansion parameter} It is of some practical importance to realize that the choice of the expansion parameter can have a considerable effect on the range over which perturbation theory gives a good approximation. Let us imagine that we introduce a new parameter $\epsilon'$ which is a function of $\epsilon$ such that $ d\epsilon'/d\epsilon$ approaches unity as $\epsilon\rightarrow0$. If we take $\epsilon'$ to be the basis of our perturbation approach, the approximation becomes \begin{eqnarray} g(x^\alpha;\epsilon)&=g(x^\alpha;\epsilon(\epsilon')) = g_{\mu\nu}(x^\alpha;\epsilon(\epsilon'))|_{\epsilon'=0}+ \epsilon'\frac{\partial}{\partial\epsilon'} g_{\mu\nu}(x^\alpha;\epsilon(\epsilon'))|_{\epsilon'=0}+{\cal O}(\epsilon'^2)\nonumber\\ &= g_{\mu\nu}(x^\alpha;\epsilon(\epsilon'))|_{\epsilon=0}+ \left[\epsilon\frac{\partial}{\partial\epsilon'} g_{\mu\nu}(x^\alpha;\epsilon(\epsilon'))|_{\epsilon=0}\right] \left\{\frac{\epsilon'}{\epsilon}\right\}+{\cal O}(\epsilon'^2)\ .\label{prime} \end{eqnarray} At $\epsilon=0$ the derivative of $g_{\mu\nu}$ with respect to $\epsilon$ and with respect to $\epsilon'$ have the same values, so for a given spacetime --- that is, for a given value of $\epsilon$ --- the nonspherical perturbation in (\ref{prime}) differs from that in (\ref{pertbasic}) by the factor $\left\{\epsilon'/\epsilon\right\}$. Computed energies (which are quadratic in the nonspherical perturbations) will differ by the square of this ratio. Different choices of parameterization will change this factor and affect the accuracy of the linearized approximation. To show the effects of this parameterization dependence, we take as an example Misner data \cite{price_pullin94}\cite{misner} for two holes. The initial separation of the holes, in units of the mass of the spacetime, is described by Misner's parameter $\mu_0$. The metric perturbations, however, are not analytic in $\mu_0$ as $\mu_0\rightarrow0$, so $\mu_0$ cannot be used as the expansion parameter in (\ref{pertbasic}). The actual expansion parameter used by Price and Pullin, was a function of $\mu_0$ denoted $\kappa_2$. We consider here what would be the results of perturbation theory done with the expansion parameter \begin{equation}\label{eq.diffk} \epsilon=\frac{\kappa_2}{1-k\kappa_2}\ . \end{equation} Figure 2 shows the results, along with the energies computed by numerical relativity applied to full nonlinear evolution \cite{anninos_etal93}. For all choices of $k$ the agreement between perturbation theory and numerical relativity is good at sufficiently small initial separation (sufficiently small $\mu_0$), but as $\mu_0$ grows larger, the agreement increasingly depends on the which parameterization is used. The $k=0$ parameterization, the parameter of the Price-Pullin paper, is a reasonably good approximation even up to separations ($\mu_0>1.36$) for which the initial apparent horizon consists of two disjoint parts. For positive values of $k$ the agreement is less impressive, while for $k=-4$, it appears that perturbation theory is giving excellent answers for initial data that are very nonspherical. Clearly the $k=-4$ parameterization is ``better,'' at least for the purpose of computing radiated energy. There exist yet better choices; in principle a parameterization could be found for which the energy computed by linearized theory is perfect for any initial separation. The crucial point is that we have no {\em a priori} way of choosing what is and what is not a good parameterization. The choice of expansion parameter $\kappa_2$ was made in the Price-Pullin analysis, because it occurred naturally in the mathematical expressions for the initial geometry. There was no {\em a priori} reason for believing it to be a particularly good, or particularly bad parameterization. This point will be discussed again, in connection with numerical results presented in Sec.~IV. The fact, demonstrated in Fig.~2, that the results of linear perturbation theory are arbitrary may seem to suggest that perturbation answers, from a formal expansion or numerical initial data, are of little value. It should be realized, however, that the arbitrariness exhibited in Fig.~2 is simply a demonstration of the fact that linearized perturbation results are uncertain to second order in the expansion parameter. The fact that the results for different parameterizations start to differ from each other around $\mu_0\approx1.5$ simply signals that $\kappa_2$ is around unity. (In fact, $\kappa_2\approx0.24$ for $\mu_0=1.2$.) Higher order uncertainty is an unavoidable feature in the range where the expansion parameter is of order unity. But there is a potential misunderstanding about the meaning of ``expansion parameter around unity.'' To see this consider a change to a new expansion parameter $\epsilon=10^{-4}*\kappa_2$. The new expansion parameter $\epsilon$ is of order unity for $\mu_0\approx7$, yet we know that perturbation fails dramatically for such a large value of $\mu_0$. The issue here is that we need some way of ascribing an appropriate ``normalization'' to the expansion parameter. A sign that the normalization is good is that physically-based measures of distortion start getting large for $\epsilon$ around unity. If we had reliable measures of this type then we could have some confidence about the range of the the expansion parameter for which we could neglect second order uncertainty, whether due to parameter arbitrariness or the omission of higher order terms in the calculation. One can formulate interesting measures for the normalization of the expansion parameter, such as the extent to which the linearized initial conditions violates the exact Hamiltonian constraint \cite{suen}. Most such measures are useful only for finding a very rough normalization for $\kappa_2$ (equivalently, for roughly finding the range in which linearized perturbation theory is reliable). The only reliable procedure for this is to carry out computations of radiated waveforms and energy to second order in the expansion parameter. The ratio of second order corrections to first order results gives the only direct measure of the reliability of perturbation results. If one computes an energy for which the second order correction to the first order result is 10\%, then one knows that the third order correction (due to a change in parameterization or an inclusion of third order terms in the computation) will be on the order of 1\%. \subsection{Treating nonlinear initial data as a perturbation expansion} We turn now to the central question of this paper: How does one apply perturbation theory to numerically generated initial data? To do this we consider our numerical initial data to be initial data for a solution in a parameterized family $g_{\mu\nu}(x^\alpha;\epsilon)$ corresponding to $\epsilon=\epsilon_{\rm num}$. The application of perturbation theory is equivalent to replacing $g_{\mu\nu}(x^\alpha;\epsilon_{\rm num})$ by \begin{equation}\label{eq.linnum} g_{\mu\nu}(x^\alpha;\epsilon)|_{\epsilon=0}+ \frac{\partial}{\partial\epsilon} g_{\mu\nu}(x^\alpha;\epsilon)|_{\epsilon=0}\epsilon_{\rm num}. \end{equation} An added familiar complication is that we can introduce a family of coordinate transformations $x^\alpha=x^\alpha(x^{\mu'};\epsilon)$ which reduces to $x^\alpha =x^{\alpha'}$ for $\epsilon\rightarrow0$. Such a transformation takes the original family to a new family $g^{'}_{\mu\nu}(x^{\alpha'};\epsilon)$, which satisfies the same requirements as the original family. We follow Moncrief\cite{moncrief74} in constructing, from the 3-metric $\gamma_{ij}$ on constant-$t$ surfaces, quantities $q_i$, which are invariant to first order in $\epsilon$ (``gauge invariant''), for coordinate transformations. The construction of these Moncrief $q_i$ is done in two steps. First, the multipole moments of the metric are extracted. In practice this is done by multiplying the metric functions by certain angular factors and integrating over angles. Since we are only interested in quadrupole and higher order for radiation, this step also eliminates the spherically symmetric background parts of the metric function. The second step is to form linear combinations of these multipoles and of their derivatives with respect to radius. We symbolically represent the process of forming these quantities as \begin{equation}\label{monconstr} q_i=Q_i(\gamma_{ij}, \partial_r \gamma_{ij})\ . \end{equation} Here the symbol ``$Q_i$'' represents the process of multiplying by angular functions and integrating, then multiplying by certain functions of $r$ and taking linear combinations of the results. (Our notation here disagrees with that of Moncrief\cite{moncrief74} in a potentially confusing way. Moncrief's perturbation quantities are independent of the size of $\epsilon$. In order to have definitions that can be applied to numerical data we use quantities that -- to first order -- are proportional to $\epsilon$.) The Moncrief gauge invariants play two different roles. For even parity one of the gauge invariants, $q_2$, is a constraint; it vanishes in linearized theory as a result of the initial value equations. In linearized theory, the remaining Moncrief quantities, denoted $q_1$ here, satisfy wave equations $L(q_1)=0$, the Regge-Wheeler equation in odd parity and Zerilli equation in even parity. {}From our numerical data we construct the quantities $q_i$ precisely according to (\ref{monconstr}). Our numerically constructed ``perturbation'' quantities will not be invariant under coordinate transformations, but rather will transform as $q'_i=q_i+{\cal O}(\epsilon_{\rm num}^2)$. Similarly, the linearized constraint, $q_2$ will not vanish, but will be of order $\epsilon_{\rm num}^2$. The numerically constructed wavefunctions $q_1$ will satisfy $L(q_1)={\cal O}((\epsilon_{\rm num})^2)$, where $L$ is the Regge-Wheeler or Zerilli wave operators. The use of NPS methods is equivalent to ignoring the second order terms in the wave equations. The wavefunction $q_i$ can then be propagated from the initial hypersurface forward and the radiation waveforms extracted from it. To evolve $q_1$ off the initial hypersurface, however, requires the initial time derivative $\partial q_1/\partial t$. This can be computed from the initial extrinsic curvature, but some care is needed. Indeed, the possible ambiguities that arise here are the justification for the somewhat protracted discussion in this section. If ${\bf n}$ is the future-directed unit normal to the initial hypersurface then the rate at which the 3-metric is changing is given by \begin{equation}\label{K=Lng} K_{ij}=-\frac{1}{2}{\cal L}_{\bf n}\,\gamma_{ij}\ , \end{equation} where $K_{ij}$ is the extrinsic curvature and ${\cal L}_{\bf n}$ is the Lie derivative along the unit normal. The unit normal is related to the derivative with respect to Schwarzschild time by $\partial/\partial t=S^{1/2}\ {\bf n}$. The time derivative of the Moncrief function then can be written \begin{displaymath} \partial q_1/\partial t=S^{1/2} {\cal L}_{\bf n}q_1 \end{displaymath}\begin{equation} =S^{1/2}{\cal L}_{\bf n}Q_1(\gamma_{ij}, \partial\gamma_{ij}/\partial r)\ . \end{equation} To evaluate the right hand side we need to know how $Q_1$ changes when it is Lie dragged by ${\bf n}$. Since $Q_1$ depends only on $\gamma_{ij}$ it might appear that one need only Lie drag $\gamma_{ij}$ to find the change in $Q_1$, and that ${\cal L}_{\bf n}Q_1=Q_1({\cal L}_{\bf n}\gamma_{ij},\partial{\cal L}_{\bf n}\gamma_{ij}/\partial r)$. From this it would follow that $\partial q_1/\partial t=-2S^{1/2} Q_1(K_{ij}, \partial K_{ij}/\partial r)$. It is important to note that this is {\em not} the correct relationship between $K_{ij}$ and the cauchy data for the wave equation. The fallacy in this procedure lies in the fact that $q_1$ must be computed from the 3-metric on a slice for which Schwarzschild time is constant (to first order in $\epsilon_{\rm num}$). Lie dragging by ${\bf n}$ moves the 3-metric to a surface that is not (to first order) a constant time surface. The cure is clearly to compare quantities on surfaces of constant $t$ by using ${\cal L}_t\equiv S^{1/2}{\cal L}_{\bf n}$. It is the Schwarzschild time derivative that commutes with the Schwarzschild radial derivative ${\cal L} _t (\partial /\partial r)^a=0$. The correct prescription then follows from \begin{displaymath} \partial q_1/\partial t=S^{1/2}{\cal L}_{\bf n}q_1 \end{displaymath}\begin{displaymath} =Q_1(S^{1/2}{\cal L}_{\bf n}\gamma_{ij}, \partial (S^{1/2}{\cal L}_{\bf n}\gamma_{ij})/\partial r) \end{displaymath} \begin{equation}\label{qdot} = -2 Q_1(S^{1/2} K_{ij}, \partial (S^{1/2}K_{ij})/\partial r)\ . \end{equation} We note that the perturbed Schwarzschild metric does have a shift vector $\beta_i$ of order $\epsilon$, and in principle the shift vector influences the time development of $\gamma_{ij}$ according to $\partial_{t} \gamma_{ij} = \partial_{t'} \gamma_{ij} + 2 \nabla_{(i}\beta_{j)}$, where $t'$ is a time coordinate in which the shift vector vanishes. But the shift vector can be considered to be ``pure gauge.'' It is necessary if one wants a complete specification of the coordinates and the metric components, but its value is a matter of choice, and is not necessary for a complete specification of the physics. The initial value, and evolution, of the gauge invariant quantity $q_1$ is invariant with respect to the choice of $\beta_i$, and $q_1$ carries all the (physically meaningful) information about gravitational waves. The evaluation of $q_1$ from (\ref{monconstr}) and $\partial q_1/\partial t$ from (\ref{qdot}) completes the extraction, from the numerical data for $\gamma_{ij}, K_{ij}$ of the cauchy data for the Regge-Wheeler or Zerilli wave equation. An alternative procedure arises if one uses the scalar wave-equations derived from the perturbative reduction of the nonlinear wave-equation for the extrinsic curvature which arises in a new explicitly hyperbolic form of the Einstein equations\cite{aacby95}. In this system, the scalar wave equations are one order lower in time derivative from the usual Regge-Wheeler and Zerilli equations, so the Cauchy data consists of the extrinsic curvature and its time-derivative (which involves the 3-dimensional Ricci curvature). {}From the above it is clear that linearized evolution should give good accuracy when applied to numerically generated initial data with sufficiently small deviations from sphericity. For initial data which are known in analytic form one can, of course, apply linearized theory even to cases in which initial deviations from sphericity are only marginally small. The results in Fig.~2, for example, show that the results of such application of perturbation theory give reasonable accuracy for values of $\mu_0$ at which an initial horizon is highly distorted. It is worrisome to apply linearized evolution to marginally nonspherical initial data, which do not, for example, satisfy the constraint $q_2=0$ with reasonable accuracy. Such a procedure --- linear evolution of nonlinear initial data --- has, among other disadvantages, no clear theoretical framework. \subsection{Calculating radiated energy by ``forced linearization''} We wish to point out here that NPS methods can be used more broadly, and a procedure we call ``forced linearization'' can be applied to numerically generated initial data in a way that amounts to extracting the linearized part of the data and evolving linearly. This procedure circumvents the difficulty of performing formal linearization to data which is only known numerically. We imagine that we start with an initial value problem in which there is some adjustable parameter, call it $\mu$, such that $\mu=0$ corresponds to the Schwarzschild initial data. There is no requirement that the family of solutions $g_{\mu\nu}(x^\alpha;\mu)$ be analytic in $\mu$ at $\mu\rightarrow0$. There may be additional parameters, call them $p_i$, such as the parameters governing the initial momenta of holes. To apply forced linearization we fix the values of the $p_i$ and make a choice of $\mu$ such that the computed initial data $\gamma^{\rm vns}_{ij},K^{\rm vns}_{ij}$ are ``very nearly spherical.'' One criterion for this would be that $q_2$ is very small. We then interpret this initial data as being essentially linearized data, to which the approximation in (\ref{eq.linnum}) applies. We extract multipoles, form a gauge invariant wave function $q_1$, and evolve it with the Zerilli or Regge-Wheeler equation, all as described above. The result of this will be a late-time waveform $q_1^{\rm vns}(r,t)$ and the energy $E^{\rm vns}$ that it carries. The next step is to characterize the results with a well behaved gauge invariant parameter. To do this we choose some fiducial radius $r_{\rm fid}$, and evaluate $\epsilon^{\rm vns}\equiv q_1(r_{\rm fid},t=0)$ the gauge invariant wave function of the initial hypersurface at this radius. Next, we leave the $p_i$ unchanged, but choose a larger value of $\mu$ for which the numerically generated initial data set $\gamma^{\rm mrgnl}_{ij},K^{\rm mrgnl}_{ij}$ is ``marginal'' in that it corresponds to deviations from sphericity large enough so that it differs significantly form linearized initial conditions; one sign of this would be that the condition $q_2=0$ is significantly violated. For this data set we go through the same procedure as above in characterizing the data set by a parameter $\epsilon^{\rm mrgnl}\equiv q_1(r_{\rm fid},t=0)$. For this marginally spherical initial data we take the solution for the wavefunction and energy to be \begin{equation} q_1^{\rm mrgnl}(r,t)=\left( \frac{\epsilon^{\rm mrgnl}}{\epsilon^{\rm vns}}\right)q_1^{\rm vns}(r,t) \ \ \ \ \ E^{\rm mrgnl}=\left( \frac{\epsilon^{\rm mrgnl}}{\epsilon^{\rm vns}}\right)^2E^{\rm vns} \ . \end{equation} The idea underlying this method is that the very nearly spherical data give us the solution for for $\partial g_{\mu\nu}/\partial\epsilon|_{\epsilon=0}$. For the marginal initial data set we then need only multiply this initial data by the appropriate factor telling us how much larger is the linear part of the nonsphericity than that of the very nearly spherical initial data. The success of forced linearization requires then that $\epsilon$ evaluated at $r_{\rm fid}$ be a well behaved parameterization of the linearized part of the nonsphericity in the numerical data. Since our expansion parameter $\epsilon$ is the magnitude of the perturbation, it will be a good expansion parameter as long as it is evaluated in a region where the nonlinear deviations from sphericity are small, i.e., where (\ref{eq.linnum}) is a good approximation. For this reason it is important that $r_{\rm fid}$ be chosen fairly large. For processes of the type pictured in Fig.~1, the deviations from sphericity fall off quickly in radius, so that at large enough $r$ one can be certain that the initial data are an excellent approximation to linearized data. Evidence for this is that the violations of the $q_2=0$ constraint are always confined to small radii. One easily implemented check on the forced linearization procedure is to look at the factor $\epsilon^{\rm mrgnl}/\epsilon^{\rm vns}$ and confirm that it is independent of $r$ for $r>r_{\rm fid}$. In Section III we show that this test is easily passed by a numerical example, and that the results of forced linearization are essentially the same as those of formal linearized theory. \section{Extraction of perturbations from numerical data} Here we assume that the reader has numerical solutions for the 3-metric on an approximately t=const surface. The first step in applying NPS to numerical results is to transform to coordinates which are ``almost Schwarzschild'' coordinates. It is assumed that the numerical $\gamma_{ij}$ and $K_{ij}$ are expressed in a coordinate system $R,\theta,\phi$ in which the approximate spherical symmetry is manifest. This means that $K_{ij}$ and ratios like \begin{equation} \frac{ \gamma_{R\theta} } {\sqrt{\gamma_{\theta\theta}}} \ \ \ \ \ \ \frac{ \gamma_{R\phi} } {\sqrt{\gamma_{\theta\theta}}} \ \ \ \ \ \frac{ \gamma_{\theta\phi} } {\gamma_{\theta\theta}} \end{equation} must be small. They all are, in fact, formally of order $\epsilon_{\rm num}$, so if they are not all reasonably small compared to unity there is little reason to think that NPS will work. A Schwarzchild-like areal radial coordinate $r$ needs to be introduced. This can be defined as a function of $R$ by \begin{equation} r\equiv\left( \int\,\gamma_{\theta\theta}\,\gamma_{\phi\phi}\,d\Omega\ \right)^{1/4}/4\pi. \end{equation} where the integral is taken on a surface of constant $R$. The metric component $\gamma_{rr}$, in terms of this quantity, gives us another test of how close the geometry is to that of a constant time Schwarzschild slice. The quantity \begin{displaymath} r\left(1-1/\gamma_{rr}\right) \end{displaymath} should be nearly equal to the constant $2M$, where $M$ is the mass of the spacetime. The variability of this quantity in $r, \theta$, and $\phi$, is formally of order $\epsilon_{\rm num}$. There are, of course, other ways of specifying the Schwarzschild-like coordinates. We could, for example, have defined $r^2\equiv\gamma_{\theta\theta}$ All these coordinate choices, however, should agree to order $\epsilon_{\rm num}$ and are therefore equivalent within a linearized gauge transformation. To compute the gauge invariant perturbation functions, we first assume that an $\ell m$ multipole of the 3-metric may be expanded as \begin{equation}\label{mncrform} \gamma_{ij}=c_1 (\hat e_1)_{ij} +c_2 (\hat e_2)_{ij} +h_1 (\hat f_1)_{ij} + {H_2 \over S} (\hat f_2)_{ij}+ r^2 K (\hat f_3)_{ij}+ +r^2 G (\hat f_4)_{ij} \end{equation} where, for clarity, we have suppressed multipole indices and have replaced Moncrief's $h_1$ and $h_2$ odd parity perturbation functions with $c_1$, $c_2$. The multipole moments $c_1, c_2, h_1, H_2, K,$ and $G$ are computed by projection onto the relevant spherical harmonics which can be found in Moncrief\cite{moncrief74}. Explicit formulas for the important special case of even parity, axisymmetric perturbations may be found in Ref.~\cite{abhss}. For odd parity perturbations, one function can be constructed from the amplitudes $c_1$ and $c_2$ which is gauge invariant and satisfies the Regge-Wheeler equation (below), \begin{equation} Q^{\times}_{\ell m} = \sqrt{2\frac{(\ell+2)!}{(\ell-2)!}}\, \left[c_1+{1 \over 2} \left ( {\partial c_2 \over \partial r} - {2 \over r} c_2 \right ) \right ] {S \over r}. \end{equation} The situation for even parity perturbations is more complicated. Two gauge invariant functions may be formed out of the multipole moments: \begin{eqnarray} k_1&=& K + {S \over r} (r^2 \partial_r G- 2 h_1) \\ k_2 &=&{1 \over 2S} \left[ H_2 - r \partial_r k_{1} - \left(1-\frac{M}{rS}\right)k_1+ S^{1/2}\partial_r (r^2 S^{1/2} \partial_r G - 2 S^{1/2} h_1) \right] \end{eqnarray} {}From $k_1$ and $k_2$ it is possible to form two new functions, one which is radiative and one which is equivalent to the perturbed hamiltonian constraint \begin{eqnarray} q_1 &=& 4 r S^2 k_2 + \ell (\ell+1) r k_1 \\ q_2 &=& \partial_r [ 4 r S^2 k_2 + \ell (\ell+1) r k_1 ] + \ell(\ell+1) [2 S k_2 + (1-M/\{rS\}) k_1 ]. \end{eqnarray} The scaled function \begin{equation} Q^+_{\ell m}={ q_1 \over \Lambda}\sqrt{2(\ell-1)(\ell+2) \over \ell(\ell+1)}\ , \end{equation} with \begin{displaymath} \Lambda\equiv(\ell-1)(\ell+2)+6M/r\ , \end{displaymath} satisfies the Zerilli equation (below). The time derivatives of the radiative gauge invariant functions $Q^{\times}_{\ell m}$ and $Q^+_{\ell m}$ are found by substituting $\sqrt{1-2M/r}K_{ij}$ for $\gamma_{ij}$ in the multipole moment computation and forming the same combinations of moments. The wavefunctions $Q^{\times}_{\ell m}$ and $Q^{+}_{\ell m}$ obey the Regge-Wheeler and Zerilli wave equations respectively: \begin{eqnarray} {\bf L} Q^{\times}_{\ell m} +V^{\times}_{\ell} Q^{\times}_{\ell m} &=& 0 \\ {\bf L} Q^{+}_{\ell m} +V^{+}_{\ell} Q^{+}_{\ell m} &=& 0 \label{eq:zer} \end{eqnarray} where the wave operator appropriate to Schwarzschild spacetime is \begin{equation} {\bf L} = {\partial ^2 \over \partial t^2} - {\partial ^2 \over \partial r_*^2} \end{equation} in terms of the ``tortoise coordinate'' $r_*=r+2M \ln (r/2M-1)$, and where the potentials are given by \begin{equation} V^{\times}_{\ell} = (1-2M/r) \left [ {\ell(\ell +1) \over r^2} -{6M \over r^3} \right ] \end{equation} and, \begin{equation} V^{+}_{\ell}(r) = (1-2M/r) \left[{1\over\Lambda^2} \left ( {72M^2\over r^5}-{12M\over r^3}(\ell-1)(\ell+2)(1-3M/r)\right) +{\ell(\ell-1)(\ell+1)(\ell+2) \over r^2 \Lambda } \right]. \end{equation} Once the Zerilli and Regge-Wheeler equations are integrated for all the desired $\ell$ and $m$ modes, the total radiated energy can be calculated from the asymptotic timeseries for $Q_{\ell m}^+$ and $Q_{\ell m}^{\times}$: \begin{equation} {d E \over dt} = {1 \over 32 \pi} \sum_{\ell=2}^\infty \sum_{m=-\ell}^\ell \left( {d Q_{\ell m}^+ \over dt}^2+ {d Q_{\ell m}^{\times} \over dt}^2 \right) {}. \label{eq:power} \end{equation} \section{Example of perturbation extraction} In this section we demonstrate the extraction of a perturbation from a numerical solution to the nonlinear constraint equations -- the Misner data representing two black holes at a moment of time symmetry. The Misner 3-geometry may be written \cite{price_pullin94} as \begin{equation} \label{eq:mismet} dl^2 = \Phi(r, \theta, \mu_0)^4(S^{-1} dr^2 + r^2 d\Omega^2) {}. \end{equation} The conformal factor $\Phi$ is given by \begin{equation} \label{eq:miscon} \Phi(r, \theta; \mu_0)= 1 + 2 (1+M/2R)^{-1}\sum_{l=2,4,...}^{\infty} \kappa_\ell \left({M \over R} \right)^{\ell+1} P_\ell(\cos \theta), \end{equation} where \begin{displaymath} R\equiv(\sqrt{r}+\sqrt{r-2M}\,)^2/4 \end{displaymath} and \begin{equation} \label{eq:miskap} \kappa_{\ell}\equiv \left({1 \over 4 \sum_{n=1}^\infty (\sinh{n \mu_0})^{-1}} \right)^{\ell+1}\, \sum_{n=1}^\infty {(\coth{n \mu_0})^\ell \over \sinh n \mu_0 }\ . \end{equation} For this exercise, we pretend that the initial geometry is known only numerically, so no explicit formal linearization can be done. The odd parity perturbations vanish in the Misner solution. We compute the even parity gauge invariant wavefunction for $\ell=2$ using numerical evaluations of (\ref{eq:miscon}) - (\ref{eq:miskap}). Specifically, we compute $K$ and $H_2$ of (\ref{mncrform}) from \begin{equation} K= H_2 = \int d \Omega \Phi^4 Y_{20}\ . \end{equation} All the other moments in (\ref{mncrform}) vanish for the conformally Schwarzschild metric of (\ref{eq:mismet}). The function $Q_{20}^+$ is evaluated at values of $r$ corresponding to the range $r_* = -20M$ to $r_*=50M$. The initial value of $Q_{20}^+$ (along with its time-derivative which is zero for the Misner time-symmetric initial data) provides initial values for integration of (\ref{eq:zer}). At large radius, $r= 100M$, the value of $\partial Q_{20}^+/\partial t$ is used in (\ref{eq:power}) to compute the radiated energy. First, in Fig.~\ref{fig.misnl} we show the result of directly computing the gauge invariant function $Q_{20}^+$ from the nonlinear initial data, integrating the Zerilli equation, and computing the radiated energy. For small values of $\mu_0$ the agreement with the explicitly linearized data of Ref.~\cite{price_pullin94} is excellent. At about $\mu_0 \simeq 1.2$ the agreement breaks down and the qualitative behavior becomes dramatically different. It is interesting to note that the apparent horizon encompassing both black holes does not exist for $\mu_0 >1.36$, close to the dramatic reversal in the energy curve. In Fig.~\ref{fig.miscon} the violation of the linearized constraint by the nonlinear data is shown as a function of radius. We plot the ratio of the constrained gauge invariant function, $q_1$ to the radiative function $q_2$ scaled in such a way as to compensate for large violation at $r=2M$. The value of $q_2$ clearly grows much faster than the radiative variable $q_1$ as the separation is increased. As discussed in Sec.~II, it is possible to obtain the results of formal perturbation theory directly from the numerical data without ever making reference to the analytic solution. In Fig.~\ref{fig.misfl} we demonstrate the application of the forced linearization procedure to the nonlinear Misner data for various values of the fiducial radius $r_{\rm fid}$. For very small values of $\mu_0$, such as $\mu_0=0.5$, the geometry outside the event horizon is everywhere well approximated by (\ref{eq.linnum}) and forced linearization works even for small values of $r_{\rm fid}/M$. When $\mu_0$ is larger than around $1.5$, on the other hand, the initial geometry near the horizon contains significant nonlinear effects, and large values of $r_{\rm fid}/M$ must be used to get results equivalent to those of formal linearized theory. As $r_{\rm fid}$ gets large, the results become indistinguishable from those of formal perturbation theory reported in Ref.~\cite{price_pullin94}. For $r_{\rm fid}=30M$ the difference in radiated energy for $\mu_0=3.0$ is less than $10^{-3} \%$. This high-accuracy equivalence deserves some explanation. In particular, why is forced linearization equivalent to formal linearization with expansion parameter $\kappa_2$? Why is that expansion parameter singled out? The equivalence is a result of two features of the way in which the linearizations were done: First, both the formal linearization of Ref.~\cite{price_pullin94}, and the forced linearization results in Fig.~\ref{fig.misfl}, use precisely the same coordinates. (The forced linearization results, in fact, are not based on initial values that were generated by genuinely numerical means. Rather, the closed form solutions for the Misner metric functions were used. The ``almost-Schwarzschild'' coordinates of the forced linearization, were precisely the same as the ``almost-Schwarzschild'' coordinates in Ref.~\cite{price_pullin94}). Secondly, in the ``almost-Schwarzschild'' coordinate system, the parameter $\kappa_2$ is, to all perturbation orders, the coefficient of the dominant nonsphericity at large radius. Forced linearization (in the limit of large $r_{\rm fid}$) results in a parameterization based on a gauge invariant measure of nonsphericity at large radius. It therefore must be proportional to $\kappa_2$ and produce results equivalent to those of the formal linearization of Ref.~\cite{price_pullin94}, in which $\kappa_2$ was the expansion parameter. It should be understood that this does not imply that the parameter $\kappa_2$ is {\em physically} singled out. A first order change in the ``almost-Schwarzschild'' coordinates will change the coefficient of the dominant large-radius nonsphericity. We might, for example, transform from the ``almost-Schwarzschild'' radial coordinate $r$ of (\ref{eq:mismet}) to a new coordinate $r' \equiv r [1 + \kappa_{2} P_2 (\cos \theta)]$. In this case the coefficient of the leading large $r'$ term in the metric will be $\kappa_2+{\cal O}(\kappa_2^2)$, and the results of forced linearization with the resulting ``numerical'' data will differ, when perturbations are large, from the results in Ref.~\cite{price_pullin94}. The forced linearization will have induced an expansion parameter different from $\kappa_2$. AMA was supported by National Science Foundation grant PHY 93-18152/ASC 93-18152 (ARPA supplemented). RHP was supported by the National Science Foundation under grants PHY9207225 and PHY9507719.
\section{Introduction} The aim of this paper is to solve, in a unified way, several mysteries which have arisen over the past few years in connection with generalizations of the notion of Lie algebra in differential geometry. T. Courant \cite{Courant:1990} introduced the following antisymmetric bracket operation on the sections of $TP\oplus T^*P$ over a manifold $P$: \[ [X_1+ \xi_1, X_2 +\xi_2]= [X_1,X_2]+ (L_{X_1}\xi_2-L_{X_2}\xi_1 + d({\textstyle{\frac 12}}(\xi_1(X_2)-\xi_2(X_1))). \] Were it not for the last term, this would be the bracket for the semidirect product of the Lie algebra ${\cal X}(P)$ of vector fields with vector space $\Omega^1(P)$ of 1-forms via the Lie derivative representation of ${\cal X}(P)$ on $\Omega^1(P)$. The last term, which was essential for Courant's work (about which more will be said later) causes the Jacobi identity to fail. Nevertheless, for subbundles $E\subseteq TP\oplus T^*P$ which are maximally isotropic for the bilinear form $(X_1 +\xi_1,X_2+ \xi_2)_+= \frac 12(\xi_1(X_2)+\xi_2(X_1))$, closure of $\Gamma(E)$ under the Courant bracket implies that the Jacobi identity {\em does} hold on $\Gamma(E)$, because of the maximal isotropic condition on $E$. These subbundles are called {\em Dirac structures} on $P$; the notion is a simultaneous generalization of that of Poisson structure (when $E$ is the graph of a map $\tilde\pi:T^*P\to TP$) and that of closed 2-form (when $E$ is the graph of a map $\tilde\omega:TP\to T^{*}P$). \begin{quote} {\bf Problem 1}. Since the Jacobi identity is satisfied on certain subspaces where $( \ , \ )_+$ vanishes, find a formula for the {\em Jacobi anomaly}\footnote{``$+ c.p.$'' below (and henceforth) will denote ``plus the other two terms obtained by circular permutations of $(1,2,3).''$} \[ [[e_1,e_2],e_3] + c.p. \] in terms of $( \ , \ )_+$. \end{quote} \bigskip The vector space $\chi(P)\oplus \Omega^1(P)$ on which the Courant bracket is defined is also a module over $C^\infty(P)$. Projection on the first factor defines a map $\rho$ from $\chi(P)\oplus \Omega^1(P)$ to derivations of $C^\infty(P)$. If one checks the Leibniz identity which enters in the definition of a {\em Lie algebroid} \cite{Mackenzie:book}, \[ [e_1,fe_2]=f[e_1,e_2]+(\rho(e_1)f)e_2 \, \] It turns out that this is not satisfied in general, but that it is satisfied for Dirac structures. This suggests: \begin{quote} {\bf Problem 2}. Express the {\em Leibniz anomaly} $[e_1,fe_2]-f[e_1,e_2]-(\rho(e_1)f)e_2$ in terms of $( \ , \ )_+$. \end{quote} When one is given an inner product on a Lie algebra, it is natural to ask whether it is invariant under the adjoint representation. Here again, a calculation turns up an {\em invariance anomaly}. We solve problems 1 and 2 in the paper, finding an expression for the invariance anomaly as well. The formulas obtained are so attractive as to suggest: \begin{quote} {\bf Problem 3}. Generalize the Courant bracket by writing down a set of axioms for a skew-symmetric bracket $\cal E\times\cal E\to\cal E$, a linear map $\cal E\to$ $\mbox{Der}(C^\infty(M))$, and a symmetric inner product $\cal E\times\cal E\to C^\infty(M)$ on the space $\cal E$ of sections of a vector bundle over $M$, and find other interesting examples of the structure thus defined. \end{quote} \bigskip Our solution of Problem 3 begins with the definition of a structure which we call a {\em Courant algebroid}{\footnote{ We apologize to our French colleagues for possible confusion with the nearly homonymous and somewhat less synonymous term, ``alg\`ebre de courants".}}. Among the examples of Courant algebroids which we find are the direct sum of any Lie bialgebroid \cite{MackenzieX:1994} and its dual, with the bracket given by a symmetrized version of Courant's original definition. This structure thus gives an answer as well to: \begin{quote} {\bf Problem 4}. What kind of object is the double of a Lie bialgebroid? \end{quote} \noindent Furthermore, within each Courant algebroid, one can consider the maximal isotropic subbundles closed under bracket. These more general Dirac structures are new Lie algebroids (and sometimes Lie bialgebroids). Constructions in this framework applied to the Lie bialgebroid of a Poisson manifold \cite{MackenzieX:1994} lead to new ways of building Poisson structures and shed new light on the theory of Poisson-Nijenhuis structures used to explicate the hamiltonian theory of completely integrable systems \cite{K-SM:1990}. In particular, we find a composition law for certain pairs of (possibly degenerate) Poisson structures which generalizes the addition of symplectic structures: namely, if $U:T^*P\to TP$ and $V:T^*P\to TP$ define Poisson structures such that $U+V$ is invertible, then $U(U+V)^{-1}V$ again defines a Poisson structure. When the base manifold $P$ is a point, a Lie algebroid is just a Lie algebra. A Courant algebroid over a point turns out to be nothing but a Lie algebra equipped with a nondegenerate ad-invariant symmetric 2-form (sometimes called an orthogonal structure \cite{me-re:algebres}). (The formulas for the anomalies all involve derivatives, so they vanish when $P$ is a point.) Such algebras and their maximal isotropic subalgebras are the ingredients of the theory of Lie bialgebras and Manin triples \cite{dr:quantum}. In fact, just as a complementary pair of isotropic subalgebras in a Lie algebra with orthogonal structure determines a Lie bialgebra, so a complementary pair of Dirac structures in a Courant algebroid determines a Lie bialgebroid. It is this fact, which exhibits our theory as a generalization of the theory of Manin triples, which is responsible for the application to Poisson-Nijenhuis pairs mentioned above. We mentioned earlier that the notion of Dirac structures was invented in order to treat in the same framework Poisson structures, which satisfy the equation $[\pi,\pi]=0$, and closed 2-forms, which satisfy $d\omega=0$. One could look for a more direct connection between these equations. \begin{quote} {\bf Problem 5}. What is the relation between the equations $[\pi,\pi]=0$ and $d\omega=0$? \end{quote} Our solution to Problem 5 is very simple. In a Courant algebroid of the form $A\oplus A^*$, the double of a Lie bialgebroid, the equation which a skew-symmetric operator $\tilde I:A\to A^*$ must satisfy in order for its graph to be a Dirac structure turns out to be the Maurer-Cartan equation $dI+\frac 12[I,I]=0$ for the corresponding bilinear form $I\in\Gamma(\wedge^2 A^*)$. The structure of the original Courant algebroid $TM\oplus TM^*$ (also viewed dually as $T^*M\oplus TM$) is sufficiently degenerate that one of the terms in the Maurer-Cartan equation drops out in each of the two cases. The next problem arises from Drinfeld's study \cite{dr:poisson} of Poisson homogeneous spaces for Poisson Lie groups. He shows in that paper that the Poisson manifolds on which a Poisson Lie group $G$ acts transitively are essentially (that is, if one deals with local rather than global objects, as did Lie in the old days) in 1-1 correspondence with Dirac subspaces of the double of the associated Lie bialgebra $({\mathfrak g} , {\mathfrak g}^{*})$. It is natural, then, to look for some kind of homogeneous space associated to a Dirac subbundle in the double of a Lie bialgebroid. The object of which a Lie bialgebroid $E\lon P$ is the infinitesimal limit is a Poisson groupoid, i.e. a Poisson manifold $\Gamma$ carrying the structure of a groupoid with base $P$, for which the graph of multiplication $\{(k,g,h)|k=gh\}$ is a coisotropic submanifold of $\Gamma \times \overline{\Gamma}\times \overline{\Gamma}$. ($\overline{\Gamma }$ is $\Gamma $ with the opposite Poisson structure. See \cite{MackenzieX:1994} \cite{MackenzieX:1996} and \cite{we:coisotropic}.) Unlike in the case of groups, a Poisson groupoid corresponding to a given Lie bialgebroid may exist only locally. \begin{quote} {\bf Problem 6}. Define a notion of Poisson homogeneous space for a Poisson groupoid. Show that Dirac structures in the double of a Lie bialgebroid $(A,A^{*})$ correspond to (local) Poisson homogeneous spaces for the (local) Poisson groupoid $\Gamma $ associated to $(A,A^{*})$. \end{quote} Our solution to Problem 6 will be contained in a sequel to this paper \cite{LWX}. Even if we work locally, it is somewhat complicated, since the ``homogeneous spaces'' for groupoids, which are already hard to define in general (see \cite{br-da-ha:topological}), in this case can involve the quotient spaces of manifolds by arbitrary foliations. To give a flavor of our results, we mention here one example. For the standard Lie bialgebroid $(TM,T^{*}M)$, the associated Poisson groupoid is the pair groupoid $M\times M$ with the zero Poisson structure. A Dirac structure transverse to $T^{*}M$ is the graph of a closed 2-form $\omega $ on $M$. The corresponding Poisson homogeneous space for $M\times M$ is $M\times (M/{\cal F})$, where the factor $M$ has the zero Poisson structure, and the factor $M/{\cal F}$ is the (symplectic) Poisson manifold obtained from reduction of $M$ by the characteristic foliation ${\cal F}$ of $\omega $. (Of course, the leaf space $M/{\cal F}$ might not be a manifold in any nice sense.) Dually, our Dirac structure also defines a Poisson homogeneous space for the Poisson groupoid of the Lie bialgebroid $(T^{*}M,TM)$, which is $T^{*}M$ with the operation of addition in fibres and the Poisson structure given by the canonical 2-form. The homogeneous space is again $T^{*}M$, with the Poisson structure coming from the sum of the canonical 2-form and the pullback of $\omega $ by the projection $T^{*}M\lon M$. \bigskip We turn now to some problems which remain unsolved. The only examples of Courant algebroids which we have given are the doubles of Lie bialgebroids, i.e. those admitting a direct sum decomposition into Dirac subbundles. For Courant algebroids over a point, there are many examples which are not of this type, even when the symmetric form has signature zero, which is necessary for such a decomposition. For instance, we may take the direct sum of two Lie algebras of dimension $k$ with invariant bilinear forms, one positive definite and one negative definite. Any isotropic subalgebra of dimension $k$ must be the graph of an orthogonal isomorphism from one algebra to the other. Such an isomorphism may not exist. Even if it does, it might be the case that the graphs of any two such isomorphisms must have a line in common. (For instance, take two copies of ${\mathfrak su}(2)$ and use the fact that every rotation of ${\mathbb R}^{3}$ has an axis.) These examples and a further study of Manin triples from the point of view of Lie algebras with orthogonal structure may be found in \cite{me-re:lie}. \begin{quote} {\bf Open Problem 1}. Find interesting examples of Courant algebroids which are not doubles of Lie bialgebroids, including examples which admit one Dirac subbundle, but not a pair of transverse ones. Are there Courant algebroids which are not closely related to finite dimensional Lie algebras, for which the bilinear form is positive definite? \end{quote} In his study of quantum groups and the Knizhnik-Zamolodchikov equation, Drinfeld \cite{dr:quasi} introduced quasi-Hopf algebras, in which the axiom of coassociativity is weakened, and their classical limits, the Lie quasi-bialgebras. The latter notion was studied in depth by Kosmann-Schwarzbach \cite{ko:quasi} (see also \cite{ba-ko:double}), who defined various structures involving a pair of spaces in duality carrying skew symmetric brackets whose Jacobi anomalies appear as coboundaries of other objects. Her structures are not subsumed by ours, though, since our expression for the Jacobi anomaly is zero when the base manifold is a point. Jacobi anomalies as coboundaries also appear in the theory of ``strongly homotopy Lie algebras'' \cite{la-ma:strongly} and in recent work of Ginzburg \cite{gi:resolution}. The relation of these studies to Courant algebroids is the subject of work in progress with Dmitry Roytenberg. \begin{quote} {\bf Open Problem 2}. Define an interesting type of structure which includes both the Courant algebroids and the Lie quasi-bialgebras as special cases. \end{quote} At the very beginning of our study, we found that if the bracket on a Courant algebroid is modified by the addition of a symmetric term, many of the anomalies for the resulting asymmetric bracket become zero. This resembles the ``twisting'' phenomenon of Drinfeld \cite{dr:quasi}. \begin{quote} {\bf Open Problem 3}. What is the geometric meaning of such asymmetric brackets, satisfying most of the axioms of a Lie algebroid? \end{quote} The next problem is somewhat vague. The Maurer-Cartan equation $d\alpha+\frac 12[\alpha,\alpha]$ appears as an integrability condition in the theory of connections and plays an essential role in modern deformation theory. (See \cite{mi:rational} and various original sources cited therein.) \begin{quote} {\bf Open Problem 4}. Find geometric or deformation-theoretic interpretations of the Maurer-Cartan equation for Dirac structures. \end{quote} Lie algebras, Lie algebroids and (the doubles of) Lie bialgebras are the infinitesimal objects corresponding to Lie groups, Lie groupoids, and (the doubles of) Poisson Lie groups respectively. Moreover, Kosmann-Schwarzbach \cite{ko:quasi} has studied the global objects corresponding to Lie quasi-bialgebras, and Bangoura \cite{ba:quasi} has recently identified the dual objects. Yet the following problem is unsolved, even for $TM\oplus T^*M$. \begin{quote} {\bf Open Problem 5}. What is the global, groupoid-like object corresponding to a Courant algebroid? In particular, what is the double of a Poisson groupoid? \end{quote} A solution to Open Problem 4 might come from a solution to the next problem. When one passes from an object such as a Lie bialgebra or even a Lie quasi-bialgebra to its double, the resulting object is frequently ``nicer'' in the sense that some of the anomalies possessed by the original object now vanish. \begin{quote} {\bf Open Problem 6}. What is the double of a Courant algebroid? \end{quote} Finally, we would like to remark that many of the constructions in this paper can be carried out at a more abstract level, either replacing the sections of a vector bundle $E$ by a more general module over $C^\infty(P)$, as in \cite{hu:poisson}, or in the context of local functionals on mapping spaces as in \cite{Dorfman} by Dorfman. \bigskip \noindent {\bf Acknowledgements.} In addition to the funding sources mentioned in the first footnote, we like to thank several institutions for their hospitality while work on this project was being done: the Isaac Newton Institute (Weinstein, Xu); the Nankai Institute for Mathematics (Liu, Weinstein, Xu); Peking University (Xu). Thanks go also to Yvette Kosmann-Schwarzbach, Jiang-hua Lu, Kirill Mackenzie and Jim Stasheff for their helpful comments. \section{Double of Lie bialgebroids} \begin{defi} \label{def:quasi-algebroid} A {\em Courant algebroid} is a vector bundle $E\lon P$ equipped with a nondegenerate symmetric bilinear form $( \cdot , \cdot )$ on the bundle, a skew-symmetric bracket $[\cdot , \cdot ]$ on $\gm (E)$ and a bundle map $\rho :E\lon TP$ such that the following properties are satisfied: \begin{enumerate} \item For any $e_{1}, e_{2}, e_{3}\in \gm (E)$, $[[e_{1}, e_{2}], e_{3}]+c.p.={\cal D} T(e_{1}, e_{2}, e_{3});$ \item for any $e_{1}, e_{2} \in \gm (E)$, $\rho [e_{1}, e_{2}]=[\rho e_{1}, \rho e_{2}];$ \item for any $e_{1}, e_{2} \in \gm (E)$ and $f\in C^{\infty} (P)$, $[e_{1}, fe_{2}]=f[e_{1}, e_{2}]+(\rho (e_{1})f)e_{2}- (e_{1}, e_{2}){\cal D} f ;$ \item $\rho \mbox{\tiny{$\circ $}} {\cal D} =0$, i.e., for any $f, g\in C^{\infty}(P)$, $({\cal D} f, {\cal D} g)=0$; \item for any $e, h_{1}, h_{2} \in \gm (E)$, $\rho (e) (h_{1}, h_{2})=([e , h_{1}]+{\cal D} (e ,h_{1}) , h_{2})+(h_{1}, [e , h_{2}]+{\cal D} (e ,h_{2}) )$, \end{enumerate} where $T(e_{1}, e_{2}, e_{3})$ is the function on the base $P$ defined by: \begin{equation} \label{eq:T0} T(e_{1}, e_{2}, e_{3})=\frac{1}{3} ([e_{1}, e_{2} ], e_{3})+c.p., \end{equation} and ${\cal D} : C^{\infty}(P)\lon \gm (E)$ is the map defined\footnote{In this paper, $d_{0}$ denotes the usual differential from functions to 1-forms, while $d$ will denote the differential from functions to sections of the dual of a Lie algebroid.} by ${\cal D} = \frac{1}{2} \beta^{-1}\rho^{*} d_{0}$, where $\beta $ is the isomorphism between $E$ and $E^*$ given by the bilinear form. In other words, \begin{equation} \label{eq:D} ({\cal D} f , e)= \frac{1}{2} \rho (e) f . \end{equation} \end{defi} {\bf Remark.} Introduce a twisted bracket (not antisymmetric!) on $\gm (E)$ by $$[e ,h \tilde{]}=[e ,h]+{\cal D} (e, h). $$ Then (iii) is equivalent to \begin{equation} [e_{1}, fe_{2} \tilde{]}=f [e_{1},e_{2}\tilde{]}+ (\rho (e_{1})f)e_{2}; \end{equation} (v) is equivalent to \begin{equation} \rho (e) (h_{1} , h_{2})=([e ,h_{1}\tilde{]}, h_{2})+ (h_{1} , [e, h_{2}\tilde{]} ); \end{equation} and (ii) and (iv) can be combined into a single equation: \begin{equation} \rho [e_{1} , e_{2}\tilde{]}=[\rho e_{1}, \rho e_{2}]. \end{equation} It would be nice to interpret equation (i) in terms of this twisted bracket. The geometric meaning of this twisted bracket remains a mystery to us. \begin{defi} Let $E$ be a Courant algebroid. A subbundle $L$ of $E$ is called {\em isotropic} if it is isotropic under the symmetric bilinear form $( \cdot , \cdot )$. It is called {\em integrable} if $\gm (L)$ is closed under the bracket $[\cdot , \cdot ]$. A {\em Dirac structure}, or {\em Dirac subbundle}, is a subbundle $L$ which is maximally isotropic and integrable. \end{defi} The following proposition follows immediately from the definition. \begin{pro} \label{prop} Suppose that $L$ is an integrable isotropic subbundle of a Courant algebroid \\ $(E, \rho , [\cdot , \cdot ], ( \cdot , \cdot )) $. Then $(L, \rho |_{L} , [\cdot , \cdot ])$ is a Lie algebroid. \end{pro} Suppose now that both $A$ and $A^{*}$ are Lie algebroids over the base manifold $P$, with anchors $a$ and $a_{*}$ respectively. Let $E$ denote their vector bundle direct sum: $E=A\oplus A^{*}$. On $E$, there exist two natural nondegenerate bilinear forms, one symmetric and another antisymmetric, which are defined as follows: \begin{equation} \label{eq:pairing} (X_{1}+\xi_{1} , X_{2}+\xi_{2})_{\pm}=\frac{1}{2} (\langle \xi_{1}, X_{2} \rangle \pm \langle \xi_{2} , X_{1}\rangle ). \end{equation} On $\gm (E)$, we introduce a bracket by \begin{equation} \label{eq:double} [e_{1}, e_{2}]=([X_{1}, X_{2}]+L_{\xi_{1}}X_{2}-L_{\xi_{2}}X_{1}-d_{*}(e_{1}, e_{2})_{-}) + ([\xi_{1} , \xi_{2}]+L_{X_{1}}\xi_{2}-L_{X_{2}}\xi_{1} +d(e_{1}, e_{ 2})_{-}), \end{equation} where $e_{1}=X_{1}+\xi_{1}$ and $e_{2}=X_{2}+\xi_{2}$.\\\\ Finally, we let $\rho : E\lon TP$ be the bundle map defined by $\rho =a +a_{*}$. That is, \begin{equation} \rho (X+\xi )=a(X)+a_{*} (\xi ) , \ \ \forall X\in \gm (A) \mbox{ and } \xi \in \gm (A^{*}) \end{equation} It is easy to see that in this case the operator ${\cal D}$ as defined by Equation (\ref{eq:D}) is given by $${\cal D}=d_{*}+d, $$ where $d_{*}: C^{\infty}(P)\lon \gm (A)$ and $d: C^{\infty}(P)\lon \gm ( A^{*})$ are the usual differential operators associated to Lie algebroids \cite{MackenzieX:1994}. When $(A, A^{*})$ is a Lie bialgebra $({\mathfrak g} , {\mathfrak g}^{*})$, the bracket above reduces to the famous Lie bracket of Manin on the double ${\mathfrak g} \oplus {\mathfrak g}^{*}$. On the other hand, if $A$ is the tangent bundle Lie algebroid $TM$ and $A^{*}=T^{*}M $ with zero bracket, then Equation (\ref{eq:double}) takes the form: $$[X_{1}+\xi_{1} , X_{2}+\xi_{2}]=[X_{1} ,X_{2}] +\{ L_{X_{1}}\xi_{2} -L_{X_{2}}\xi_{1} +d(e_{1}, e_{ 2})_{-} \}.$$ This is the bracket first introduced by Courant \cite{Courant:1990}, then generalized to the context of the formal variational calculus by Dorfman \cite{Dorfman}. Our work in this paper is largely motivated by an attempt to unify the two examples above, based on the observation that Courant's bracket appears to be some kind of ``double.'' In order to generalize Manin's construction to Lie algebroids, it is necessary to have a compatibility condition between Lie algebroid structures on a vector bundle and its dual. Such a condition, providing a definition of {\em Lie bialgebroid}, was given in \cite{MackenzieX:1994}. We quote here an equivalent formulation from \cite{K-S:1994}. \begin{defi} A {\em Lie bialgebroid} is a dual pair $(A,A^*)$ of vector bundles equipped with Lie algebroid structures such that the differential $d_*$ on $\Gamma(\wedge^*A)$ coming from the structure on $A^*$ is a derivation of the Schouten-type bracket on $\Gamma(\wedge^*A)$ obtained by extension of the structure on $A$. \end{defi} The following two main theorems of this paper show that we have indeed found the proper version of the theory of Manin triples for the Lie algebroid case. \begin{thm} \label{thm:main1} If $(A, A^{*})$ is a Lie bialgebroid, then $E=A \oplus A^{*}$ together with $([\cdot , \cdot ], \rho , (\cdot , \cdot)_{+})$ is a Courant algebroid. \end{thm} Conversely, we have \begin{thm} \label{thm:main2} In a Courant algebroid $(E, \rho , [\cdot , \cdot ], ( \cdot , \cdot ))$, suppose that $L_{1}$ and $L_{2}$ are Dirac subbundles transversal to each other, i.e., $E=L_{1}\oplus L_{2}$. Then, $(L_{1}, L_{2} )$ is a Lie bialgebroid, where $L_{2}$ is considered as the dual bundle of $L_{1}$ under the pairing $2( \cdot , \cdot )$. \end{thm} An immediate consequence of the theorems above is the following duality property of Lie bialgebroids, which was first proved in \cite{MackenzieX:1994} and then by Kosmann-Schwarzbach \cite{K-S:1994} using a simpler method. \begin{cor} If $(A, A^{*})$ is a Lie bialgebroid, so is $(A^{*}, A)$. \end{cor} \section{Jacobi anomaly} In this section, we begin the computations leading to the proofs of our main theorems. Throughout this section, we assume that $A$ is a Lie algebroid with anchor $a$ and that its dual $A^*$ is also equipped with a Lie algebroid structure with anchor $a_{*}$. However, we shall {\em not} assume any compatibility conditions between these two algebroid structures. For simplicity, for any $e_{i}=X_{i}+\xi_{i}\in \gm (E), i=1,2,3$, we let $$J(e_{1}, e_{2}, e_{3})=[[e_{1}, e_{2} ], e_{3}]+c.p.$$ The main theorem of this section is the following \begin{thm} \label{thm:main-jacobi} Assume that both $(A, a)$ and $(A^{*}, a_{*})$ are Lie algebroids. Then, for $e_{i}=X_{i}+\xi_{i}\in \gm (E), i=1,2,3$, we have \begin{equation} J(e_{1}, e_{2}, e_{3})={\cal D} T( e_{1}, e_{2}, e_{3})-(J_{1} +J_{2} +c.p ), \end{equation} \end{thm} where $$J_{1} =i_{X_{3}}(d [\xi_{1} , \xi_{2} ]-L_{\xi_{1}}d \xi_{2} +L_{\xi_{2}}d \xi_{1} ) + i_{\xi_{3}} (d_{*} [X_{1}, X_{2}]-L_{X_{1}}d_{*}X_{2} +L_{X_{2}}d_{*}X_{1} ), $$ and $$J_{2} = L_{d_{*}(e_{1}, e_{2})_{-}}\xi_{3} + [d(e_{1}, e_{2})_{-},\ \xi_{3}] +L_{d (e_{1}, e_{2})_{-}}X_{3}+[d_{*}(e_{1}, e_{2})_{-},\ X_{3} ]. $$ We need a series of lemmas before proving this theorem. \begin{lem} For $e_{i}=X_{i}+\xi_{i}\in \gm (E), i=1,2,3$, $T$ is skew-symmetric, and \begin{equation} \label{eq:T} T(e_{1}, e_{2}, e_{3})=\frac{1}{2} \{\langle [X_{1}, X_{2}], \xi_{3} \rangle +\langle [\xi_{1}, \xi_{2}], X_{3}\rangle +a(X_{3})(e_{1}, e_{2})_{-}-a_{*}(\xi_{3})(e_{1}, e_{2})_{-}\}+c.p. \end{equation} \end{lem} \noindent{\bf Proof.}\ The first assertion is obvious from the definition of $T$. For the second one, we first have \be &&([e_{1}, e_{2}], e_{3})_{+}\\ &=&\frac{1}{2} \{ \langle [X_{1}, X_{2}], \xi_{3} \rangle + \langle L_{\xi_{1}}X_{2}, \xi_{3}\rangle - \langle L_{\xi_{2}}X_{1} , \xi_{3}\rangle -a_{*}(\xi_{3}) (e_{1}, e_{2})_{-}\\ &&+\langle [\xi_{1}, \xi_{2}], X_{3}\rangle +\langle L_{X_{1}}\xi_{2} ,X_{3} \rangle -\langle L_{X_{2}}\xi_{1} , X_{3}\rangle +a(X_{3})(e_{1}, e_{2})_{-}\}\\ &=&\frac{1}{2} \{[\langle [X_{1}, X_{2}], \xi_{3} \rangle+\langle [\xi_{1}, \xi_{2}], X_{3}\rangle ]+c.p. \}\\ &&+\frac{1}{2} \{a_{*}(\xi_{1})\langle X_{2} ,\xi_{3} \rangle -a_{*}(\xi_{2})\langle X_{1}, \xi_{3} \rangle -a_{*}(\xi_{3})(e_{1}, e_{2})_{-}\\ &&\ \ \ \ \ \ \ \ +a(X_{1})\langle \xi_{2} ,X_{3}\rangle -a(X_{2})\langle \xi_{1} ,X_{3}\rangle +a(X_{3})(e_{1}, e_{2})_{-} \}\\ &=&\frac{1}{2} [\{\langle [X_{1}, X_{2}], \xi_{3} \rangle +\langle [\xi_{1}, \xi_{2}], X_{3}\rangle +a(X_{3})(e_{1}, e_{2})_{-}-a_{*}(\xi_{3})(e_{1}, e_{2})_{-}\}+c.p.]\\ &&+\frac{1}{2} \rho (e_{1}) (e_{2}, e_{3})_{+}-\frac{1}{2} \rho (e_{2}) (e_{3}, e_{1})_{+} \ee Therefore, by taking the sum of its cyclic permutations, one obtains \be T(e_{1}, e_{2}, e_{3})&=&\frac{1}{3} ([e_{1}, e_{2}], e_{3})_{+}+c.p.\\ &=& \frac{1}{2} \{\langle [X_{1}, X_{2}], \xi_{3} \rangle +\langle [\xi_{1}, \xi_{2}], X_{3}\rangle +a(X_{3})(e_{1}, e_{2})_{-}-a_{*}(\xi_{3})(e_{1}, e_{2})_{-}\}+c.p. \ee \qed As a by-product, we obtain the following identity by substituting Equation (\ref{eq:T}) into the last step of the computation of $([e_{1}, e_{2}], e_{3})_{+}$ in the proof above. This formula will be useful later. \begin{equation} \label{eq:T1} ([e_{1}, e_{2}], e_{3})_{+}=T(e_{1}, e_{2}, e_{3}) +\frac{1}{2} \rho (e_{1}) (e_{2}, e_{3})_{+}-\frac{1}{2} \rho (e_{2}) (e_{3}, e_{1})_{+} \end{equation} \begin{lem} \label{lem:4.3} \begin{equation} i_{X}L_{\xi}d\eta =[\xi, L_{X}\eta ]-L_{L_{\xi}X} \eta +[d\langle \eta , X\rangle ,\ \xi ] +d (a_{*}(\xi)\langle \eta , X\rangle ) -d\langle [\xi , \eta ] ,X\rangle . \end{equation} \end{lem} \noindent{\bf Proof.}\ For any $Y\in \gm (A)$, \be \langle i_{X}L_{\xi} d \eta , Y\rangle &=& (L_{\xi }d\eta )(X, Y)\\ &=& a_{*}(\xi )[d\eta (X, Y)]-d\eta (L_{\xi }X, Y)- d\eta (X, L_{\xi }Y)\\ &=&a_{*}(\xi ) a(X) \langle \eta , Y\rangle -a_{*}(\xi )a(Y)\langle \eta , X\rangle -a_{*}(\xi )\langle \eta , [X, Y]\rangle \\ &&-a(L_{\xi }X )\langle \eta , Y\rangle +a(Y)\langle \eta , L_{\xi}X\rangle +\langle \eta , [L_{\xi} X , Y] \rangle \\ &&-a(X) \langle \eta , L_{\xi }Y\rangle +a(L_{\xi }Y)\langle \eta , X\rangle +\langle \eta , [X, L_{\xi}Y]\rangle\\ &=&a_{*}(\xi )\langle L_{X}\eta , Y\rangle -a_{*}(\xi )a(Y)\langle \eta , X\rangle +a(Y)a_{*}(\xi ) \langle \eta , X\rangle\\ &&-a(Y)\langle [\xi , \eta ], X\rangle -\langle L_{L_{\xi}X}\eta , Y\rangle -\langle L_{X}\eta , L_{\xi }Y\rangle +\langle L_{\xi}Y, d\langle \eta ,X\rangle \rangle \\ &=&\langle [\xi , L_{X}\eta ], Y\rangle +\langle [d \langle \eta , X\rangle ,\xi ], Y\rangle \\ &&+a(Y)a_{*}(\xi )\langle \eta , X\rangle -a(Y) \langle [\xi ,\eta ], X\rangle -\langle L_{L_{\xi }X}\eta , Y\rangle. \ee The lemma follows immediately. \qed \begin{lem} \label{lem:T1} \begin{equation} \label{eq:14} ([e_{1}, e_{2}], e_{3})_{-}+c.p.=T(e_{1}, e_{2}, e_{3})+ [ \{a(X_{3})(e_{1}, e_{2})_{-}+2a_{*}(\xi_{3})(e_{1}, e_{2})_{-} - \langle [\xi_{1}, \xi_{2}], X_{3}\rangle \} +c.p. ] \end{equation} \end{lem} \noindent{\bf Proof.}\ By definition, $$([e_{1}, e_{2}], e_{3})_{-}+([e_{1}, e_{2}], e_{3})_{+} =\langle [e_{1}, e_{2}]^{*}, X_{3}\rangle, $$ where $[e_{1}, e_{2}]^{*}$ refers to the component of $[e_{1}, e_{2}]$ in $\gm (A^{*})$. It thus follows that \be &&\{([e_{1}, e_{2}], e_{3})_{-}+ c.p. \}+ 3T(e_{1}, e_{2}, e_{3})\\ &=& \langle [e_{1}, e_{2}]^{*} , X_{3} \rangle +c.p.\\ &=& \langle [\xi_{1}, \xi_{2}]+L_{X_{1}}\xi_{2}-L_{X_{2}}\xi_{1} +d(e_{1}, e_{ 2})_{-}, \ \ X_{3} \rangle +c.p.\\ &=& \{\langle [\xi_{1}, \xi_{2}], X_{3}\rangle + a(X_{1})\langle \xi_{2}, X_{3}\rangle -\langle \xi_{2} , [X_{1}, X_{3}]\rangle\\ &&+a(X_{3})(e_{1}, e_{2})_{-}-a(X_{2})\langle \xi_{1} , X_{3} \rangle +\langle \xi_{1} , [X_{2}, X_{3}]\rangle \} +c.p.\\ &=& \{\langle [\xi_{1}, \xi_{2}], X_{3}\rangle +2 \langle [X_{1}, X_{2}], \xi_{3}\rangle +3a(X_{3})(e_{1}, e_{2})_{-} \}+c.p.\\ &=&4T(e_{1}, e_{2}, e_{3})+[\{a(X_{3})(e_{1}, e_{2})_{-} +2a_{*}(\xi_{3})(e_{1}, e_{2})_{-}-\langle [\xi_{1}, \xi_{2}], X_{3}\rangle \}+c.p. ], \ee where the second from the last step follows essentially from reorganizing cyclic permutation terms and the last step uses Equation (\ref{eq:T}). Equation (\ref{eq:14}) thus follows immediately. \qed {\bf Proof of Theorem \ref{thm:main-jacobi}} We denote by $I_{1}$ and $I_{2}$ the components of $J(e_{1}, e_{2}, e_{3})$ on $\gm (A^{*})$ and $\gm (A)$ respectively. Thus, by definition, \be I_{1} &=&\{ [[\xi_{1}, \xi_{2}], \xi_{3}]+[L_{X_{1}}\xi_{2}-L_{X_{2}}\xi_{1} ,\xi_{3}]+[d(e_{1}, e_{2})_{-}, \xi_{3}]\\ &&\ \ +L_{ [X_{1}, X_{2}]+L_{\xi_{1}}X_{2}-L_{\xi_{2}}X_{1}-d_{*}(e_{1}, e_{ 2})_{-} }\xi_{3}\\ &&\ \ \ \ +L_{X_{3}}L_{X_{2}}\xi_{1}-L_{X_{3}}L_{X_{1}}\xi_{2}-L_{X_{3}}[\xi_{1}, \xi_{2}]\\ &&\ \ \ \ \ \ -d[ a(X_{3})(e_{1}, e_{ 2})_{-}] +d([e_{1}, e_{2}], e_{3})_{-} \}+ c.p. \ee By using the Jacobi identity: $[[\xi_{1}, \xi_{2}], \xi_{3}]+c.p. =0$ and the relation: $L_{[X_{1}, X_{2}]}=[L_{X_{1}}, L_{X_{2}}]$, we can write \begin{eqnarray} I_{1}&=&\{[L_{X_{1}}\xi_{2}-L_{X_{2}}\xi_{1} ,\xi_{ 3}]+ L_{L_{\xi_{1}}X_{2}-L_{\xi_{2}}X_{1}}\xi_{3} \nonumber \\ &&-L_{d_{*}(e_{1}, e_{2})_{-}}\xi_{3} + [d(e_{1}, e_{2})_{-},\ \xi_{3}]-L_{X_{3}}[\xi_{1}, \xi_{2}] \nonumber\\ &&+d([e_{1}, e_{2}], e_{3})_{-}-d(a(X_{3})(e_{1}, e_{2})_{-})\}+c.p. \label{eq:11} \end{eqnarray} Now, \be L_{X_{3}}[\xi_{1}, \xi_{2}] &=& (di_{X_{3}}+i_{X_{3}}d )[\xi_{1}, \xi_{2}]\\ &=&d\langle X_{3}, [\xi_{1}, \xi_{2}] \rangle + i_{X_{3}}L_{\xi_{1}}d\xi_{2}-i_{X_{3}}L_{\xi_{2}}d\xi_{1}\\ &&\ \ \ \ \ \ +i_{X_{3}}(d[\xi_{1}, \xi_{2}]-L_{\xi_{1}}d\xi_{2}+L_{\xi_{2}}d\xi_{1}). \ee By using Lemma \ref{lem:4.3} twice, we can write \begin{eqnarray} &&L_{X_{3}}[\xi_{1}, \xi_{2}]+c.p. \nonumber \\ &=&\{ [L_{X_{1}}\xi_{2}-L_{X_{2}}\xi_{1}, \xi_{3}]+ L_{L_{\xi_{1}}X_{2}-L_{\xi_{2}}X_{1}}\xi_{3}+2[d(e_{1}, e_{2})_{-},\ \xi_{3}] \nonumber \\ &&+2d( a_{*}(\xi_{3} )(e_{1}, e_{2})_{-}) -d\langle X_{3}, [\xi_{1}, \xi_{2}] \rangle +i_{X_{3}}(d[\xi_{1}, \xi_{2}]-L_{\xi_{1}}d\xi_{2}+L_{\xi_{2}}d\xi_{1})\}+c.p. \label{eq:12} \end{eqnarray} Substituting Equation (\ref{eq:12}) into Equation (\ref{eq:11}), we have $$I_{1}=\{ d [ ([e_{1}, e_{2}], e_{3})_{-}-a(X_{3})(e_{1}, e_{2})_{-} -2a_{*}(\xi_{3})(e_{1}, e_{2})_{-} +\langle [\xi_{1}, \xi_{2}], X_{3} \rangle ] -K_{1} -K_{2} \}+c.p., $$ where $$ K_{1} =i_{X_{3}}(d [\xi_{1} , \xi_{2} ]-L_{\xi_{1}}d \xi_{2} +L_{\xi_{2}}d \xi_{1} )$$ and $$K_{2} = L_{d_{*}(e_{1}, e_{2})_{-}}\xi_{3} + [d(e_{1}, e_{2})_{-},\ \xi_{3}].$$ It follows from Lemma \ref{lem:T1} that $$I_{1}=dT( e_{1}, e_{2}, e_{3})-\{K_{1} +K_{2} +c.p.\}. $$ A similar formula for $I_{2}$ can be obtained in a similar way. The conclusion follows immediately. \qed \section{Proof of Theorem 2.5} This section is devoted to the proof of Theorem \ref{thm:main1}. Throughout the section, we assume that $(A, A^{*})$ is a Lie bialgebroid and $E=A\oplus A^{*}$ as in Theorem \ref{thm:main1}. We also let ${\cal D}: C^{\infty}(P)\lon \gm (E)$ and $\rho :E\lon TP$ be defined as in Section 2. To prove Theorem \ref{thm:main1}, it suffices to verify all the five identities in Definition \ref{def:quasi-algebroid}. First, Equation (i) follows directly from Theorem \ref{thm:main-jacobi} and properties of Lie bialgebroids. Equation (iv) is equivalent to saying that $aa_{*}^{*}$ is skew symmetric, which is again a property of a Lie bialgebroid (see Proposition 3.6 in \cite{MackenzieX:1994}). Below, we shall split the rest of the proof into several propositions. \begin{pro} For any $f\in C^{\infty}(P)$ and $e_{1}, e_{2}\in \gm (E)$, we have \begin{equation} [e_{1}, fe_{2}]=f[e_{1}, e_{2}]+(\rho (e_{1})f)e_{2}- (e_{1}, e_{2})_{+} {\cal D} f . \end{equation} \end{pro} \noindent{\bf Proof.}\ Suppose that $e_{1}=X_{1}+\xi_{1}$ and $e_{2}=X_{2}+\xi_{2}$. Then, we have $$[e_{1}, fe_{2}]=[X_{1}, fX_{2}]+[X_{1}, f\xi_{2}]+[\xi_{1}, fX_{2}] +[\xi_{1} , f\xi_{2}].$$ Here, we have \be &&[X_{1}, fX_{2}] =f [X_{1},X_{2}]+(a(X_{1})f)X_{2};\\ &&[ \xi_{1} , f\xi_{2}]=f[\xi_{1}, \xi_{2}]+((a_{*}\xi_{1} )f)\xi_{2};\\ &&[X_{1}, f\xi_{2}]=f[X_{1}, \xi_{2}]+((aX_{1})f)\xi_{2} -\frac{1}{2} \langle X_{1}, \xi_{2} \rangle {\cal D} f ;\\ &&[\xi_{1}, fX_{2}]=f[\xi_{1},X_{2}] +((a_{*}\xi_{1} ) f)X_{2}- \frac{1}{2} \langle X_{2} , \xi_{1} \rangle {\cal D} f . \ee The conclusion follows from adding up all the equations above. \qed \begin{pro} \label{pro:hom} For any $e_{1}, e_{2}\in \gm (E)$, we have $$\rho [e_{1}, e_{2} ]=[\rho e_{1}, \rho e_{2}]. $$ \end{pro} We need a lemma first. \begin{lem} \label{lem:4.2} If $(A, A^{*})$ is a Lie bialgebroid with anchors $(a, a_{*})$, then for any $X\in \gm (A) $ and $\xi \in \gm (A^{*})$, $$ [a(X), a_{*}(\xi )] = a_{*}(L_{X}\xi )-a (L_{\xi }X) + a a_{*}^{*}d_{0} \langle \xi , X \rangle . $$ \end{lem} \noindent{\bf Proof.}\ For any $f\in C^{\infty}(P)$, \be &&(a a_{*}^{*}d_{0} \langle \xi , X \rangle ) f\\ &=& \langle d_{*}\langle \xi , X \rangle , df \rangle\\ &=&L_{df}\langle \xi , X \rangle\\ &=&\langle L_{df}\xi, X \rangle +\langle \xi, L_{df} X \rangle\\ &=& -\langle L_{\xi} df , X\rangle + \langle \xi, [X, d_{*}f]\rangle\\ &=& -a_{*}(\xi )\langle df , X\rangle +\langle df , L_{\xi} X \rangle + a(X)a_{*}(\xi )f-\langle L_{X}\xi , d_{*}f\rangle\\ &=&[a(X) , a_{*}(\xi )]f-a_{*}(L_{X}\xi )f+ a(L_{\xi} X )f, \ee where in the fourth equality we have used the fact that $L_{df}X=[X , d_{*}f]$, a property of a general Lie bialgebroid (see Proposition 3.4 of \cite{MackenzieX:1994}). \qed {\bf Proof of Proposition \ref{pro:hom}} Let $e_{1}=X_{1}+\xi_{1}$ and $e_{2}=X_{2}+\xi_{2}$. \be \rho [e_{1}, e_{2}]&=& a\{ [X_{1}, X_{2}]+L_{\xi_{1}}X_{2}-L_{\xi_{2}}X_{1}- d_{*}(e_{1}, e_{2} )_{-}\}\\ &&+a_{*}\{[\xi_{1} ,\xi_{2} ]+L_{X_{1}}\xi_{2}-L_{X_{2}}\xi_{1} + d(e_{1}, e_{2})_{-}\}\\ &=&a[X_{1}, X_{2}]+a(L_{\xi_{1}}X_{2})-a(L_{\xi_{2}}X_{1})- \frac{1}{2} a a_{*}^{*}d_{0} (\langle \xi_{1}, X_{2}\rangle- \langle\xi_{2}, X_{1} \rangle )\\ &&+a_{*}[ \xi_{1} , \xi_{2} ]+a_{*}(L_{X_{1}}\xi_{2})-a_{*}(L_{X_{2}}\xi_{1}) +\frac{1}{2} a_{*}a^{*} d_{0} (\langle \xi_{1}, X_{2}\rangle -\langle \xi_{2}, X_{1} \rangle )\\ &=&a[X_{1}, X_{2}]+[a(L_{\xi_{1}}X_{2})-a_{*}(L_{X_{2}}\xi_{1}) -aa_{*}^{*} d_{0} \langle \xi_{1}, X_{2} \rangle ]\\ &&-[ a(L_{\xi_{2}}X_{1})-a_{*}(L_{X_{1}}\xi_{2})-aa_{*}^{*}d_{0} \langle \xi_{2} , X_{1} \rangle ] +a_{*}[\xi_{1}, \xi_{2}]\\ &=&[aX_{1}, aX_{2}]+[a_{*}\xi_{1}, a_{*}\xi_{2}] +[aX_{1}, a_{*}\xi_{2}]+[a_{*}\xi_{1}, aX_{2}]\\ &=&[\rho (e_{1}) , \rho (e_{2})], \ee where in the third equality we have used the skew-symmetry of the operator $aa_{*}^{*}$, and the second from the last follows from Lemma \ref{lem:4.2}. \qed \begin{pro} For any $e, h_{1}, h_{2}\in \gm (E)$, we have \begin{equation} \label{eq:17} \rho (e) (h_{1}, h_{2})_{+}=([e , h_{1}]+{\cal D} (e ,h_{1})_{+} , \ h_{2})_{+}+ (h_{1}, \ [e , h_{2}]+{\cal D} (e ,h_{2})_{+} )_{+} \end{equation} \end{pro} \noindent{\bf Proof.}\ According to Equation (\ref{eq:T1}), $$([e ,h_{1}], h_{2})_{+}=T(e, h_{1}, h_{2})+\frac{1}{2} \rho(e) (h_{1}, h_{2})_{+} -\frac{1}{2} \rho (h_{1})(e, h_{2})_{+}$$ and $$(h_{1}, [e, h_{2}])_{+}=T(e, h_{2}, h_{1})+\frac{1}{2} \rho(e) (h_{2}, h_{1})_{+} -\frac{1}{2} \rho ( h_{2}) (e, h_{1})_{+}.$$ By adding these two equations, we obtain Equation (\ref{eq:17}) immediately since $T(e, h_{1}, h_{2})$ is skew-symmetric with respect to $h_{1}$ and $h_{2}$. \qed \section{Proof of Theorem 2.6} This section is devoted to the proof of Theorem \ref{thm:main2}. We denote sections of $L_{1}$ by letters $X, Y$, and sections of $L_{2}$ by $\xi , \eta $ etc.. For any $X\in \gm (L_{1})$ and $\xi \in \gm (L_{2})$, we define their pairing by \begin{equation} \langle \xi , X \rangle =2( \xi , X ). \end{equation} Since $(\cdot , \cdot )$ is nondegenerate, $L_{2}$ can be considered as the dual bundle of $L_{1}$ under this pairing. Moreover, the symmetric bilinear form $(\cdot , \cdot )_{+} $ on $E$ defined by Equation (\ref{eq:pairing}) coincides with the original one. By Proposition \ref{prop}, both $L_{1}$ and $L_{2}$ are Lie algebroids, and their anchors are given by $a=\rho |_{L_{1}}$ and $a_{*}=\rho |_{L_{2}}$ respectively. We shall use $d: \gm (\wedge^{*}L_{2})\lon \gm (\wedge^{*+1} L_{2} ) $ and $d_{*}: \gm (\wedge^{*}L_{1})\lon \gm (\wedge^{*+1} L_{1} )$ to denote their induced de-Rham differentials as usual. Equation (v) in Definition \ref{def:quasi-algebroid} implies immediately that the bracket between $ X\in \gm (L_{1}) $ and $ \xi \in \gm (L_{2}) $ is given by \begin{equation} [X ,\xi ]= (-L_{\xi } X +\frac{1}{2} d_{*} \langle \xi , X \rangle ) +( L_{X}\xi -\frac{1}{2} d\langle \xi , X \rangle ) \end{equation} Thus we have \begin{pro} Under the decomposition $E=L_{1}\oplus L_{2}$, for sections $e_{i}\in \gm (E)$, $i=1, 2$ if we write $e_{i}=X_{i}+\xi_{i}$, then the bracket $[e_{1}, e_{2}]$ is given by Equation (\ref{eq:double}). \end{pro} Before proving Theorem \ref{thm:main2}, we need the following lemma. \begin{lem} \label{lem:exchange} Under the assumption of Theorem \ref{thm:main2} we have \be &&L_{d_{*}f}\xi =-[df , \xi]; \\ &&L_{df}X=-[d_{*}f , X], \ee for any $f\in C^{\infty}(P)$, $X\in \gm (L_{1})$ and $\xi\in \gm (L_{2})$. \end{lem} \noindent{\bf Proof.}\ Clearly, Equation (iv) in Definition \ref{def:quasi-algebroid} yields that $a\mbox{\tiny{$\circ $}} d_{*}=-a_{*} \mbox{\tiny{$\circ $}} d$. Therefore, \begin{eqnarray} [a_{*} \xi , a X]&=&[\rho \xi , \rho X ] \nonumber\\ &=&\rho [\xi , X] \nonumber \\ &=&\rho ( L_{\xi } X -\frac{1}{2} d_{*} \langle \xi , X \rangle -L_{X}\xi +\frac{1}{2} d\langle \xi , X \rangle ) \nonumber\\ &= &a( L_{\xi } X -\frac{1}{2} d_{*} \langle \xi , X\rangle )+ a_{*}(-L_{X}\xi + \frac{1}{2} d\langle \xi , X \rangle) \nonumber\\ &=&a( L_{\xi } X)-a_{*}(L_{X}\xi )+ ( a_{*} d ) \langle \xi , X \rangle . \label{eq:15} \end{eqnarray} On the other hand, \begin{eqnarray} ((a_{*}d )\langle \xi , X \rangle ) f &=& (a (d_{*}f))\langle \xi ,X \rangle \nonumber \\ &=&\langle L_{d_{*}f}\xi , X\rangle +\langle \xi , [d_{*}f , X]\rangle \nonumber \\ &=&\langle [\xi , df], X \rangle -\langle \xi , L_{X}d_{*}f\rangle +\langle L_{d_{*}f}\xi +[df , \xi ], X\rangle \nonumber \\ &=&a_{*}(\xi )a (X)f- \langle df , L_{\xi }X\rangle - a(X)a_{*}(\xi )f +\langle L_{X}\xi , d_{*}f\rangle +\langle L_{d_{*}f}\xi +[df , \xi ], X\rangle \nonumber \\ &=&[a_{*}(\xi ), a(X)]f-a(L_{\xi }X)f+a_{*} (L_{X}\xi )f +\langle L_{d_{*}f}\xi +[df , \xi ], X\rangle . \label{eq:16} \end{eqnarray} Comparing Equation (\ref{eq:15}) with ( \ref{eq:16}), we obtain $$\langle L_{d_{*}f}\xi +[df , \xi ] , X\rangle =0.$$ Therefore, $L_{d_{*}f}\xi =-[df , \xi ]$. The other equation can be proved similarly. \qed {\bf Proof of Theorem \ref{thm:main2} } It follows from Theorem \ref{thm:main-jacobi} that $J_{1} +J_{2} +c.p. =0$, for any $e_{1}, e_{2}$ and $e_{3}\in \gm (E)$. Using Lemma \ref{lem:exchange}, we have $J_{1} + c.p. =0$. In particular, if we take $e_{1}=X_{1}, \ e_{2}=X_{2}$ and $e_{3}=\xi_{3}$, we obtain that $i_{\xi_{3}} (d_{*} [X_{1}, X_{2}]-L_{X_{1}}d_{*}X_{2} +L_{X_{2}}d_{*}X_{1} )=0$, which implies the compatibility condition between $A$ and $A^{*}$. \qed \section{Hamiltonian operators} Throughout this section, we will assume that $(A, A^{*})$ is a Lie bialgebroid. Suppose that $H: A^{*}\lon A$ is a bundle map. We denote by $A_{H}$ the graph of $H$, considered as a subbundle of $E=A\oplus A^{*}$. I.e., $A_{H}=\{ H\xi +\xi |\xi\in A^{*}\}$. \begin{thm} $A_{H}$ is a Dirac subbundle iff $H$ is skew-symmetric and satisfies the following Maurer-Cartan type equation: \begin{equation} \label{eq:Maure} d_{*}H+\frac{1}{2} [H, H]=0. \end{equation} In this equation, $H$ is considered as a section of $\wedge^{2}A$. \end{thm} In the sequel, we shall use the same symbol to denote a section of $\wedge^{2}A$ and its induced bundle map if no confusion is caused. \\ \noindent{\bf Proof.}\ It is easy to see that $A_{H}$ is isotropic iff $H$ is skew-symmetric. For any $\xi , \eta \in \gm (A^{*})$, let \begin{equation} [\xi , \eta ]_{H}=L_{H\xi}\eta -L_{H\eta }\xi +d \langle \xi , H\eta \rangle. \end{equation} Since \be &&[H\xi , \eta ]=L_{H\xi }\eta -L_{\eta }H\xi - \frac{1}{2} (d-d_{*})\langle \eta , H\xi \rangle \ \ \mbox{and} \\ &&[\xi , H\eta ]=L_{\xi}H\eta -L_{H\eta }\xi +\frac{1}{2} (d-d_{*})\langle \xi , H\eta \rangle , \ee then, $$[H\xi , \eta ]+ [\xi , H\eta ]=[\xi , \eta ]_{H}+L_{\xi}H\eta -L_{\eta}H\xi +d_{*}\langle \eta , H\xi \rangle .$$ On the other hand, we have the following formula (see \cite{K-SM:1990}): \begin{equation} [H\xi , H\eta ]=H[\xi , \eta ]_{H}-\frac{1}{2} [H , H](\xi ,\eta ). \end{equation} Therefore, we have, \be [H\xi +\xi , H\eta +\eta ]&=&[H\xi , H\eta ]+[\xi , H\eta ]+[H\xi , \eta ]+ [\xi , \eta ]\\ &=&(L_{\xi}H\eta -L_{\eta}H\xi +d_{*}\langle \eta , H\xi \rangle +H[\xi , \eta ]_{H}-\frac{1}{2} [H, H](\xi , \eta ))\\ && +([\xi , \eta ]+[\xi , \eta ]_{H} ). \ee It thus follows that $A_{H}$ is integrable iff for any $\xi , \eta \in \gm (A^{*})$ \begin{equation} \label{eq:25} H[\xi , \eta ]= L_{\xi}H\eta -L_{\eta}H\xi +d_{*}\langle \eta , H\xi \rangle -\frac{1}{2} [H, H](\xi , \eta ). \end{equation} On the other hand, \be (d_{*}H ) (\xi , \eta , \zeta )&=&a_{*}(\xi )\langle \eta , H\zeta \rangle -a_{*}(\eta )\langle \xi , H\zeta \rangle +a_{*}(\zeta )\langle \xi , H\eta \rangle\\ &&-\langle [\xi ,\eta ], H\zeta \rangle +\langle [\xi , \zeta ], H\eta \rangle -\langle [\eta , \zeta ], H\xi \rangle\\ &=&\langle H[\xi , \eta ]+L_{\eta }H\xi -L_{\xi }H\eta +d_{*}\langle \xi , H\eta \rangle , \zeta \rangle . \ee Hence, \begin{equation} (d_{*}H ) (\xi , \eta )=H[\xi ,\eta ]+L_{\eta }H\xi -L_{\xi }H\eta -d_{*}\langle \eta , H\xi \rangle . \end{equation} This implies that Equation (\ref{eq:25}) is equivalent to $$(d_{*}H )(\xi , \eta ) +\frac{1}{2} [H, H](\xi , \eta )=0, $$ or equivalently $$d_{*}H+\frac{1}{2} [H, H]=0. $$ \qed {\bf Remark} (1). Because of the symmetric role of $A$ and $A^*$, we have the following assertion: the graph $A_{I}=\{ X+I X|X\in \gm ( A )\}$ of a bundle map $I: A\lon A^{*}$ defines a Dirac subbundle iff $I$ is skew-symmetric and satisfies the following Maurer-Cartan type equation: \begin{equation} dI+\frac{1}{2} [I, I]=0. \end{equation} (2). For the canonical Lie bialgebroid $(TM, T^{*}M)$ where $M$ is equipped with the zero Poisson structure, Equation (\ref{eq:Maure}) becomes $[H,H]=0$, which is the defining equation for a Poisson structure. On the other hand, if we exchange $TM$ and $T^{*}M$, and consider the Lie bialgebroid $(T^{*}M , TM )$, the bracket term drops out of Equation (\ref{eq:Maure}), whose solutions correspond to a presymplectic structures. Encompassing these two cases into a general framework was indeed the main motivation for Courant \cite{Courant:1990} to define and study Dirac structures.\\ (3). The Maurer-Cartan equation is a kind of integrability equation. It is also basic in deformation theory, where it may live on a variety of differential graded Lie algebras. It would be interesting to place the occurrence of this equation in our theory in a more general context. \begin{defi} Given a Lie bialgebroid $(A, A^{*})$, a section $H\in \gm (\wedge^{2} A)$ is called a {\em hamiltonian operator } if $A_{H}$ defines a Dirac structure. $H$ is called a {\em strong hamiltonian operator} if $A_{\lambda H}$ are Dirac subbundles for all $\lambda \in {\mathbb R}$. \end{defi} \begin{cor} For a Lie bialgebroid $(A, A^{*})$, $H\in \gm (\wedge^{2} A)$ is a hamiltonian operator if Equation (\ref{eq:Maure}) holds. It is a strong hamiltonian operator if $d_{*}H=[H, H]=0. $ \end{cor} For a hamiltonian operator $H$, $A_{H}$ is a Dirac subbundle which is transversal to $A$ in $A\oplus A^{*}$. Therefore, $(A, A_{H})$ is a Lie bialgebroid according to Theorem \ref{thm:main2}. In fact, $A_{H}$ is isomorphic to $A^{*}$, as a vector bundle, with the anchor and bracket of its Lie algebroid structure given respectively by $\hat{a}_{*}=a_{*}+a\mbox{\tiny{$\circ $}} H$ and $[\xi , \eta \hat{]}=[\xi , \eta ]+[\xi , \eta]_{H}$, for all $\xi , \eta \in \gm (A^{*})$. In particular, if $H$ is a strong hamiltonian operator, one obtains a one parameter family of Lie bialgebroids transversal to $A$, which can be considered as a deformation of the Lie bialgebroid $(A, A^{*})$. Conversely, any Dirac subbundle transversal to $A$ corresponds to a hamiltonian operator in an obvious way. For example, consider the standard Lie bialgebra $(\mathfrak{k}, \mathfrak{b})$ arising from the Iwasawa decomposition of $\mathfrak{k}^{{\mathbb C}}$ \cite{LuW}, where $\mathfrak{k} $ is a compact semi-simple Lie algebra and $\mathfrak{b}$ its corresponding dual Lie algebra. Then any real form of $\mathfrak{k}^{{\mathbb C}}$ which is transversal to $\mathfrak{b}$ will correspond to a hamiltonian operator (see \cite{LiuQ} for a complete list of such real forms for simple Lie algebras). It is straightforward to check that such a hamiltonian operator is not strong. On the other hand, for the Cartan subalgebra $\mathfrak{h}$ of $\mathfrak{k}$, every element in $\wedge^{2}h$ gives rise to a strong hamiltonian operator $H: \mathfrak{k}^{*} \lon \mathfrak{k}$, This gives rise to a deformation of the standard Lie bialgebra $(\mathfrak{k}, \mathfrak{b})$ (see \cite{LS}). These examples can be generalized to any gauge algebroid associated to a principal $K$-bundle. Below, we will look at hamiltonian operators in two special cases, each of which corresponds to some familiar objects. \begin{ex} Let $P$ be a Poisson manifold with Poisson tensor $\pi$. Let $(TP, T^* P)$ be the canonical Lie bialgebroid associated to the Poisson manifold $P$ and $E=TP \oplus T^* P$ equipped with the induced Courant algebroid structure. It is easy to see that a bivector field $H$ is a hamiltonian operator iff $H+\pi$ is a Poisson tensor. $H$ is a strong hamiltonian operator iff $H$ is a Poisson tensor Schouten-commuting with $\pi$. \end{ex} \begin{ex} Similarly, we may switch $TP$ and $T^{*}P$, and consider the Lie bialgebroid $(T^{*}P, TP )$ associated to a Poisson manifold $P$ with Poisson structure $\pi$. Let $E= T^{*}P \oplus TP$ be equipped with its Courant algebroid structure. In this case, a hamiltonian operator corresponds to a two-form $\omega \in \Omega^{2}(P)$ satisfying $d\omega + \frac{1}{2} [\omega , \omega ]_{\pi }=0$. Here, $[\cdot , \cdot ]_{\pi}$ refers to the Schouten bracket of differential forms on $P$ induced by the Poisson structure $\pi$. Given a hamiltonian operator $\omega$, its graph $A_{\omega }$ defines a Dirac subbundle transversal to $T^{*}P$, the first component of $E$ also being considered as a Dirac subbundle. Therefore, they form a Lie bialgebroid. Their induced Poisson structure on the base space can be easily checked to be given by $-2(\pi^{\#}+N\pi^{\#}) $, where $N: TP \lon TP$ is the composition $\pi^{\#}\mbox{\tiny{$\circ $}} \omega^{b}$ and $\omega^{b} : TP \lon T^{*}P$ is the bundle map induced by the two-form $\omega$. If $\omega $ is a strong hamiltonian operator, then $N \pi^{\#} $ defines a Poisson structure compatible with $\pi$. In fact, in this case, $(\pi , N)$ is a Poisson-Nijenhuis structure in the sense of \cite{K-SM:1990}. \end{ex} We note that Vaisman \cite{va:complementary} has studied 2-forms on Poisson manifolds satisfying the condition $[\omega , \omega ]_{\pi}=0$. Such forms, called {\em complementary} to the Poisson structure, also give rise to new Lie algebroid structures on $TM$. To end this section, we describe a example of Lie bialgebroids, where both the algebroid and its dual arise from hamiltonian operators. \begin{pro} Let $U$ and $V$ be Poisson tensors over a manifold $M$ and denote by $T^{*}M_{U}$ and $T^{*}M_{V}$ their associated canonical cotangent Lie algebroids on $T^{*}M$. Assume that $U-V$ is nondegenerate. Then $(T^{*}M_{U}, T^{*}M_{V})$ is a Lie bialgebroid, where their pairing is given by $(\xi , \eta )=(U-V)(\xi , \eta )$ for any $\xi \in T^{*}M_{U}$ and $\eta \in T^{*}M_{V}$. Their induced Poisson tensor on the base space $M$ is given by $-2U(U-V)^{-1}V$. \end{pro} \noindent{\bf Proof.}\ Let $E=TM\oplus T^* M$ be equipped with the usual Courant bracket. Since both $U$ and $V$ are Poisson tensors, their graphs $A_{U}$ and $A_{V}$ are Dirac subbundles. They are transversal since $U-V$ is nondegenerate. Therefore, $(A_{U}, A_{V})$ is a Lie bialgebroid, where their pairing is given by \begin{equation} \langle\!\langle U\eta +\eta , V\xi +\xi \rangle\!\rangle = \frac{1}{2} \langle \xi , (U-V)\eta \rangle . \end{equation} On the other hand, as Lie algebroids, $A_{U}$ and $ A_{V}$ are clearly isomorphic to cotangent Lie algebroids $T^{*}M_{U}$ and $ T^{*}M_{V}$ respectively. Moreover, their anchors $a_{U}: A_{U}\lon TM $ and $a_{V}: A_{V}\lon TM $ are given respectively by $a_{U} (U\xi +\xi )=U\xi $ and $a_{V} (V\xi +\xi )=V\xi $. This proves the first part of the proposition. To calculate their induced Poisson structure on the base $M$, we need to find out the dual map $a_{V}^{*}: T^{*}M\lon A_{V}^{*}\cong A_{U}$. For any $\xi \in T^{*}M$, we assume that $a_{V}^{*} (\xi )=U\eta +\eta \in A_{U}$ via the identification above. For any $\zeta \in T^{*}M$, \be ( a_{V}^{*} \xi , V\zeta +\zeta )&=& \langle \xi , a_{V} (V\zeta +\zeta )\rangle \\ &=& \langle \xi , V\zeta \rangle \ee On the other hand, $( a_{V}^{*} \xi , V\zeta +\zeta )= \langle\!\langle U\eta +\eta , V\zeta +\zeta \rangle\!\rangle =\frac{1}{2} \langle \zeta , (U-V)\eta \rangle$. It thus follows that $\eta =-2(U-V)^{-1}V\xi $. Therefore, according to Proposition 3.6 in \cite{MackenzieX:1994}, the induced Poisson structure $a_{U}\mbox{\tiny{$\circ $}} a_{V}^{*}: T^{*}M\lon TM$ is given by $(a_{U}\mbox{\tiny{$\circ $}} a_{V}^{*})\xi =-2U(U-V)^{-1}V\xi $. \qed Replacing $V$ by $-V$ in the proposition above, we obtain the following ``composition law'' for Poisson structures. \begin{cor} \label{cor-composition} Let $U$ and $V$ be Poisson tensors over manifold $M$ such that $U+V$ is nondegenerate. Then, $U(U+V)^{-1}V$ also defines a Poisson tensor on $M$. \end{cor} Note that, if $U$ and $V$ are nondegenerate, then $U(U+V)^{-1}V=(U^{-1}+V^{-1})^{-1}$ is the Poisson tensor corresponding to the sum of the symplectic forms for $U$ and $V$. Since the sum of closed forms is closed, it is obvious in this case that $U(U+V)^{-1}V$ is a Poisson tensor. We do not know such a simple proof of Corollary \ref{cor-composition} in the general case. \section{Null Dirac structures and Poisson reduction} In this section, we consider another class of Dirac structures related to Poisson reduction and dual pairs of Poisson manifolds. \begin{pro} \label{pro:null} Let $(A, A^{*} )$ be a Lie bialgebroid, and $h \subseteq A$ a subbundle of $A$. Then $L=h\oplus h^{\perp} \subseteq E=A\oplus A^* $ is a Dirac structure iff $h$ and $h^{\perp}$ are, respectively, Lie subalgebroids of $A$ and $A^* $. \end{pro} \noindent{\bf Proof.}\ Obviously, $L=h\oplus h^{\perp}$ is a maximal isotropic subbundle of $E$. If $L$ is a Dirac structure, clearly $h$ and $h^{\perp}$ are Lie subalgebroids of $A$ and $A^*$ respectively. Conversely, suppose that both $h$ and $h^{\perp}$ are Lie subalgebroids of $A$ and $A^*$ respectively. To prove that $L$ is a Dirac structure, it suffices to show that $[X, \xi ]$ is a section of $L$ for any $X \in \gm (h) $ and $\xi \in \gm (h^{\perp})$. Now $$[X,\xi]=L_{X}\xi-L_{\xi}X .$$ For any section $Y\in \gm (h)$, $$<L_{X}\xi, \ Y>=a(X)<\xi,Y>-<\xi,[X,Y]>=0. $$ Therefore, $L_{X}\xi $ is still a section of $h^{\perp}$. Similarly, $L_{\xi}X $ is a section of $h$. This concludes the proof of the proposition. \qed It is clear that a subbundle $L\subseteq E $ is of the form $L=h\oplus h^{\perp}$ iff the minus two-form $(\cdot,\cdot)_{-}$ on $E$, as defined by Equation (\ref{eq:pairing}), vanishes on $L$. For this reason, we call a Dirac structure of this form a {\em null Dirac structure}. An immediate consequence of Proposition \ref{pro:null} is the following: \begin{cor} \label{cor:reduction} Let $(P, \ \pi )$ be a Poisson manifold, and $D$ a subbundle of $TP$. Let $T^{*}P$ be equipped with the cotangent Lie algebroid structure so that $(TP, \ T^{*}P)$ is a Lie bialgebroid. Then $L=D\oplus D^{\perp} $ is a Dirac structure in $E= TP\oplus T^{*}P$ iff $D$ is an integrable distribution and the Poisson structure on $P$ descends to a Poisson structure on the quotient space $P/D $ \footnote{When the quotient space is not a manifold, this means that at each point there is a local neighborhood $U$ such that the Poisson structure on $U$ descends to its quotient.} such that the natural projection is a Poisson map. \end{cor} \noindent{\bf Proof.}\ This follows directly from the following \begin{lem} \label{lem:reduction} Let $D$ be an integrable distribution on a Poisson manifold $P$. $P/D$ has an induced Poisson structure (in the above general sense) iff $D^{\perp}\subset T^{*}P$ is a subalgebroid\footnote{Such a foliation is also called cofoliation by Vaisman \cite{vaisman:book}}. \end{lem} \noindent{\bf Proof.}\ For simplicity, let us assume that $P/D$ is a manifold. The general case will follow from the same principle. It is clear that a function $f$ is constant along leaves of $D$ iff $df $ is a section of $D^{\perp}$. If $D^{\perp}$ is a subalgebroid, it follows from the equation \begin{equation} \label{eq:dfg} d\{f,\ g\}=[df, dg] \end{equation} that $C^{\infty}(P/D)$ is a Poisson algebra. Conversely, a local one-form $fdg$ is in $D^{\perp}$ iff $g$ is constant along $D$. The conclusion thus follows from Equation (\ref{eq:dfg}) together with the Lie algebroid axiom relating the bracket and anchor. \qed {\bf Remark.} \ Poisson reduction was considered by Marsden and Ratiu in \cite{MR}. Lemma \ref{lem:reduction} can be considered as a special case of their theorem when $P=M$ in the Poisson triple $(P, \ M, \ E)$ (see \cite{MR}). It would be interesting to interpret their general reduction theorem in terms of Dirac structures as in Corollary \ref{cor:reduction}. The rest of the section is devoted to several examples of Corollary \ref{cor:reduction}, which will lead to some familiar results in Poisson geometry. Recall that, given a Poisson Lie group $G$ and a Poisson manifold $M$, an action $$ \sigma : G\times M \lon M$$ is called a Poisson action if $\sigma$ is a Poisson map. In this case, $M$ is called a Poisson $G$-space. Now consider $P=G \times M$ with the product Poisson structure and diagonal $G$-action. Then $P/G$ is isomorphic to $M$, and the projection from $P$ to $P/G=M$ becomes the action map $\sigma$, which is a Poisson map when $P/G$ is equipped with the given Poisson structure on $M$. By Corollary \ref{cor:reduction}, we obtain a null Dirac structure $L=D\oplus D^{\perp}$ in $TP\oplus T^*P$. Clearly, $L$ is a Lie algebroid over $P$, which is $G$-invariant. It would be interesting to explore the relation between this algebroid and the one defined on $ (M\times {\mathfrak g})\oplus T^*M $, which was studied by Lu in \cite{Lu}. For a Poisson Lie group $G$ with tangential Lie bialgebra $({\mathfrak g} ,{\mathfrak g}^{*})$, the Courant algebroid double $E= TG \oplus T^{*}G$ can be identified, as a vector bundle, with the trivial product $G \times ({\mathfrak g} \oplus {\mathfrak g}^{*}) $ via left translation. Under such an identification, a left invariant null Dirac structure has the form $L=G\times (h\oplus h^{\perp})$, where $h$ is a subalgebra of ${\mathfrak g}$ and $h^{\perp}$ is a subalgebra of ${\mathfrak g}^{*}$. Thus, one obtains the following reduction theorem: for a connected closed subgroup $H$ with Lie algebra $h$, $G/H$ has an induced Poisson structure iff $h^{\perp}$ is a subalgebra of ${\mathfrak g}^*$. More generally, let $G$ be a Poisson group, $M$ a Poisson $G$-space. Suppose that $H\subseteq G$ is a closed subgroup with Lie algebra $h$. Assume that $M/H$ is a nice manifold such that the projection $p: M\lon M/H$ is a submersion. Then the $H$-orbits define an integrable distribution ${\mathfrak h}$ on $M$. According to Corollary \ref{cor:reduction}, the Poisson structure on $M$ descends to $M/H$ iff ${\mathfrak h}^{\perp}$ is a subalgebroid of the cotangent algebroid $T^{*}M$ of the Poisson manifold $M$. On the other hand, we have \begin{pro} \label{pro-perp} If $h^{\perp}$ is a subalgebra of ${\mathfrak g}^*$, then ${\mathfrak h}^{\perp}$ is a subalgebroid of $T^{*}M$. Conversely, if the isotropic subalgebra at each point is a subalgebra of $h$, and in particular if the action is locally free, then that ${\mathfrak h}^{\perp}$ is a subalgebroid implies that $h^{\perp}\subseteq {\mathfrak g}^*$ is a subalgebra. \end{pro} \noindent{\bf Proof.}\ It is easy to see that ${\mathfrak h}^{\perp}\cong \phi^{-1}(h^{\perp})$, where $\phi : T^{*}M\lon {\mathfrak g}^{*}$ is the momentum mapping for the lifted $G$-action on $T^{*}M$, equipped with the canonical cotangent symplectic structure. According to Proposition 6.1 in \cite{Xu}, $\phi : T^{*}M\lon {\mathfrak g}^{*}$ is a Lie algebroid morphism. Before continuing, we need the following \begin{lem} Let $A\lon M$ be a Lie algebroid with anchor $a$, ${\mathfrak g}$ a Lie algebra, and $\phi :A\lon {\mathfrak g}$ an algebroid morphism. Suppose that $h\subseteq {\mathfrak g}$ is a subalgebra such that $\phi^{-1}h \subseteq A$ is a subbundle. Then $\phi^{-1}h$ is a subalgebroid. Conversely, given a subalgebroid $B\subseteq A$, if $\phi (B|_{m})$ is independent of $m\in M$, then it is a subalgebra of ${\mathfrak g}$. \end{lem} \noindent{\bf Proof.}\ This follows directly from the following equation (see \cite{Mackenzie:1992}): $$ \phi [X,Y]=(aX)(\phi Y)-(aY)(\phi X)+{[\phi X, \ \phi Y]}^{.}, \ \ \forall X, Y\in \gm (A), $$ where $\phi X$, $\phi Y$ and $\phi [X, Y]$ are considered as ${\mathfrak g}$-valued functions on $P$, and $[\cdot , \cdot]^{\cdot}$ refers to the pointwise bracket. \qed Now, the first part of Proposition \ref{pro-perp} is obvious according to the lemma above. For the second part, we only need to note that $\phi ({\mathfrak h}^{\perp} ) = h^{\perp}\cap Im \phi $, and the assumption that the isotropic subalgebra at each point is a subalgebra of $h$ is equivalent to that $ h^{\perp}\subseteq Im \phi $. This concludes the proof of the Proposition. \qed From the above discussion, we have the following conclusion: if $h^{\perp}$ is a subalgebra of ${\mathfrak g}^*$, then the Poisson structure on $M$ descends to $M/H$. This is a well-known reduction theorem of Semenov-Tian-Shansky \cite{STS} (see also \cite{we:coisotropic}). Conversely, if the isotropic subalgebra at each point is a subalgebra of $h$, and in particular if the action is locally free, the converse also holds. Another interesting example arises when $P$ is a symplectic manifold with an invertible Poisson tensor $\pi$. In this case, $\pi^{\#}: \ T^{*}P\lon TP$ is a Lie algebroid isomorphism. Given a null Dirac structure $L=D\oplus D^{\perp}$, $\bar{D}= \pi^{\#}(D^{\perp})$ is a subalgebroid of $TP$. It is simple to see that $(\bar{D})^{\perp}=(\pi^{\#})^{-1}(D)$, and is therefore a subalgebroid of $T^{*}P$. Thus, $\bar{L} \stackrel{def}{=}\bar{D}\oplus (\bar{D})^{\perp}$ defines another null Dirac structure. It is easy to see that $D$ and $\bar{D}$ are symplectically orthogonal to each another. Thus $P/\bar{D}$ is a Poisson manifold (assume that it is a nice manifold) so that $P/D$ and $P/\bar{D}$ constitute a full dual pair, which is a well known result of Weinstein \cite{we:1983}. Conversely, it is clear that a full dual pair corresponds to a null Dirac structure.
\section{Introduction} \mbox{} The Haldane Shastry Model [H,S] has been a subject of much attention in the past years. This Model is a version of the XXX Heisenberg Spin Chain where spins interact with a potential inversely proportional to the squared distance between the spins. Like the XXX Chain with nearest-neighbour interaction, the Haldane Shastry Model is Integrable in the sense that its Hamiltonian is a member of a family of mutually commuting, independent Integrals of Motion. The Haldane Shastry Model, however, has been solved in much greater detail than the nearest-neighbour XXX Model. The reason for this is a remarkably large algebra of symmetries found in the former Model. For the Haldane Shastry Spin Chain with the spins taking values in the fundamental representation of ${\frak g}{\frak l}_n$ the symmetry algebra is the ${\frak g}{\frak l}_n$-Yangian. This infinite-dimensional algebra of symmetries facilitated computation of explicit expressions for the energy levels and the eigenvectors! [HHTBP,BPS,H2]. Some of the corre lation functions have been found as well [HZ]. An important reason for the attention attracted by the Haldane Shastry Model is the fractional statistics exhibited by elementary excitations over the antiferromagnetic ground state present in this Model [H2]. In the case of ${\frak g}{\frak l}_2$-Haldane Shastry Chain these excitations are spin-1/2 particles that are ``semions'' -- particles with statistics exactly half-way between bosons and fermions in the sense of Haldane. Being exactly solvable the Haldane Shastry Model provides a valuable laboratory for a study of physical implications of the fractional statistics. A remarkable connection exists between the Haldane Shastry Spin Chain in the limit of infinite number of sites and WZNW Conformal Field Theory at level 1 [BPS2,BLS]. This connection has led to a novel description of the space of states in WZNW level-1 CFT, where the states are organized into irreducible multiplets of Yangian symmetry algebra inhereted from the Haldane Shastry Model. This, in turn, provided an explanation for the Fermionic Virasoro character formulas for the level-1 integrable representations of $\hat{{\frak s}{\frak l}}_2$ that were earlier derived by [DKKMM]. In the paper [BGHP] it was realized that the Haldane Shastry Spin Chain is related to a more general class of Integrable Models -- the so-called Dynamical Models that describe quantum particles with spin moving along a circle. These Dynamical Models can be thought of as generalzations of the Calogero-Sutherland Model with inverse squared sine potential. The precise way in which the Haldane Shastry Hamiltonian and associated higher conserved charges are obtained from the hierarchy of Dynamical Models was recently explained by Polychronakos in [P], and by Talstra and Haldane in [TH]. In the last paper the authors have shown how the hierarchy of Integrable Spin Models including the Haldane Shastry Model appears in the static limit of the Dynamical Models in which the coordinates of the particles are ``frozen'' along the circle in an equidistantly spaced lattice. In the present paper we define trigonometric counterparts of the spin-1/2 Haldane Shastry Hamiltonian and the associated higher conserved charges. The hierarchy of Integrable Spin Chain Hamiltonians that we obtain has an infinite-dimensional symmetry algebra $U_q'(\hat{{\frak g}{\frak l}}_2)$ that takes over the role played by the Yangian in the Haldane Shastry Model. We compute the eigenvalue spectrum of the hierarchy and find that it has an additive, particle-like form. The space of states is decomposed into eigenspaces of the operators that form the hierarchy. Each of the eigenspaces is a highest-weight irreducible representation of $U_q'(\hat{{\frak g}{\frak l}}_2)$ parametrized by a sequence of integers --``magnon quasimomenta'' or a ``motif'' in terminology of [HHTBP,BPS]. The eigenvalue that corresponds to such eigenspace is a $q$-deformation of the eigenvalue of the Haldane Shastry Hamiltonian parametrized by the same magnon quasimomenta. The procedure that is used to derive the trigonometric hierarchy has been inspired by the Talstra-Haldane approach -- we extract the Integrable Spin Models from a static limit of the trigonometric $U_q'(\hat{{\frak g}{\frak l}}_2)$-invariant Dynamical Models that were defined in [BGHP]. The spectrum of the Spin Models is obtained by a descent from the spectrum of these Dynamical Models. \subsection{A survey of the method and results} In this subsection we highlight the main steps of the procedure that is used to define the trigonometric hierarchy of Integrable Spin Models and formulate the results of this article. The details and proofs of the statements are contained in the main body of the paper starting with sec.1. We derive the hierachy of trigonometric Spin Models by a two-step reduction from the trigonometric Dynamical Models that were introduced in [BGHP]. First of all we recall the definition of these trigonometric Dynamical Models. \subsubsection{Trigonometric Dynamical Models of [BGHP]} The trigonometric Dynamical Models are defined starting with two representations of the finite-dimensional Hecke Algebra $H_N(q)$ (Cf. {\bf 1.1}). The first of these representations is defined in the ring of polynomials in $N$ variables ${\Bbb C}[z_1,\dots,z_N]$. The generators $g_{i,i+1} \quad (i=1,\dots,N-1)$ of $H_N(q)$ in this representation have the following form \begin{equation} g_{i,i+1} :=\frac{q^{-1}z_i - qz_{i+1}}{z_i-z_{i+1}}(K_{i,i+1} - 1) + q \qquad (i=1,\dots,N-1) . \end{equation} Where $ K_{i,j} $ is the exchange operator for variables $ z_i , z_j $. \\ The second representation of $H_N(q)$ is defined in the $N$-fold tensor product of two-dimensional vector spaces $H := V^{\otimes N},\; V:= {\Bbb C}^2 = {\Bbb C}\{ v^+ , v^-\}$. The Hecke generator $t_{i,i+1} \quad (i=1,\dots,N-1)$ is a matrix acting in $H$ according to the formula \begin{equation} t_{i,i+1} = I\otimes \dots \otimes I\otimes\underset{i,i+1}{t}\otimes I \otimes \dots \otimes I \quad (i=1,\dots,N-1), \end{equation} where $t$ is the matrix which acts in $ V\otimes V$: \begin{eqnarray} t v^+\otimes v^- & = & (q-q^{-1})v^+\otimes v^- + v^-\otimes v^+ \\ t v^-\otimes v^+ & = & v^+\otimes v^- \\ t v^\pm\otimes v^\pm & = & qv^\pm\otimes v^\pm. \end{eqnarray} The two of these representations naturally extend to $ {\Bbb C}[z_1,\dots,z_N]\otimes H$. The Hecke representation generated by $g_{i,i+1} \quad (i=1,\dots,N-1)$ is enlarged to a representation of the Affine Hecke Algebra $\widehat{H_N(q)}$ by adjoining affine generators $ Y_i \quad (i=1,\dots,N)$ : \begin{eqnarray} Y_i & := & g_{i,i+1}^{-1}K_{i,i+1}\dots g_{i,N}^{-1}K_{i,N}p^{D_i}K_{1,i}g_{1,i}\dots K_{i-1,i}g_{i-1,i}. \end{eqnarray} Where $ p $ is a $c$-number , $ D_i := z_i \frac{\partial}{\partial z_i} $ and \begin{equation} g_{i,j} :=\frac{q^{-1}z_i - qz_j}{z_i-z_j}(K_{i,j} - 1) + q \qquad (i,j = 1,\dots,N-1) . \end{equation} After introducing the objects just described the authors of [BGHP] define the hierarchy of Dynamical Models. The operators $ \Delta^{(n)} \quad (n=1,\dots, N)$ that constitute the hierarchy are coefficients of the polynomial $\Delta (u) = \sum_{n=0}^N u^n \Delta^{(n)}$ which generates elementary symmetric functions of $ Y_1,Y_2,\dots,Y_N $ : \begin{equation} \Delta (u) : = \prod_{i=1}^N ( 1 + u Y_i ) . \end{equation} Due to the mutual commutativity of the operators $ Y_1,Y_2,\dots,Y_N $ the hierachy of Dynamical Models is integrable: \begin{equation} [ \Delta(u) , \Delta(v) ] = 0 . \end{equation} \mbox{} In the space $ {\Bbb C}[z_1,\dots,z_N]\otimes H $ one defines a representation of the algebra $ U_q'(\hat{{\frak g}{\frak l}}_2) $ (Cf. {\bf 1.2}). In this representation the generators of $ U_q'(\hat{{\frak g}{\frak l}}_2) $ are obtained by expanding in the parameter $u$ the monodromy matrix $ T_a(u) $ which is defined in a standard way as a product of elementary $L$-operators [BGHP]: \begin{equation} T_a(u) : = L_{a1}(uY_1)L_{a2}(uY_2)\dots L_{aN}(uY_N). \end{equation} Where the elementary $L$-operator $L_{ai}(uY_i)\quad (i=1,\dots,N)$ acts in the tensor product of an auxiliary copy of the two-dimensional vector space denoted by $V_a$ and the space $ {\Bbb C}[z_1,\dots,z_N]\otimes H$: \begin{eqnarray} L_{ai}(uY_i) & := & \frac{uY_it_{a,i} - t_{a,i}^{-1}}{uY_i - 1}P_{a,i} \quad (i=1,\dots,N). \end{eqnarray} Here $P$ is the permutation operator in $ V\otimes V$ and $ t_{a,i} \, , \, P_{a,i} $ are the usual extensions of $ t \, , \, P$ as operators in $ V_a\otimes H$. The fact that $ T_a(u) $ defines a representation of $ U_q'(\hat{{\frak g}{\frak l}}_2) $ follows from the $RTT = TTR$ relation which involves the trigonometric $R$-matrix: \begin{align} \bar{R}_{ab}(u/v)T_a(u)T_b(v) & = T_b(v)T_a(u))\bar{R}_{ab}(u/v), \\ \bar{R}(z) & := \frac{ zt - t^{-1} }{qz - q^{-1}}P. \end{align} The hierarchy of Dynamical Models defined by $\Delta (u)$ is $U_q'(\hat{{\frak g}{\frak l}}_2)$-invariant, that is \begin{equation} [ \Delta (u) , T_a(v) ] = 0 . \end{equation} This again follows from the mutual commutativity of the operators $Y_1,Y_2,\dots,Y_N.$ Both $\Delta (u)$ and $ T_a(u) $ act in the ``bosonic'' subspace of $ {\Bbb C}[z_1,\dots,z_N]\otimes H $ as explained in [BGHP]. This subspace which we denote by ${\cal B}$ is defined by the requirement of the Hecke-invariance: \begin{equation} {\cal B} := \{ b \in {\Bbb C}[z_1,\dots,z_N]\otimes H | (g_{i,i+1} - t_{i,i+1})b = 0 \quad (i=1,\dots,N-1) \}. \end{equation} In any operator $O$ acting in ${\cal B}$ one can eliminate the coordinate exchange operators $K_{i,j}$ by carrying them one-by-one to the right of any expression in $O$ and replacing a $ K_{i,j} $ standing on the right of an expression in accordance with the rule: $ g_{i,i+1} \rightarrow t_{i,i+1} \quad (i=1,\dots,N-1) $. This leads to a uniquely defined operator $\widehat{O}$ which does not contain coordinate permutations, such that \begin{equation} O{\cal B} = \widehat{O}{\cal B} . \end{equation} Where the notation means that the equality holds for any vector in ${\cal B}$.\\ With this definition the eq. (0.0.9,.12,.14) lead to \begin{align} [ \widehat{\Delta(u)} , \widehat{\Delta(v)}]{\cal B} & = 0 , \\ [ \widehat{\Delta (u)} , \widehat{T_a(v)} ]{\cal B} & = 0 , \\ (\bar{R}_{ab}(u/v)\widehat{T_a(u)}\widehat{T_b(v)} & - \widehat{T_b(v)}\widehat{T_a(u)}\bar{R}_{ab}(u/v)){\cal B} = 0 . \end{align} The set of relations (0.0.17-.19) constitutes the result of [BGHP] concerned witn the trigonometric Dynamical Models. \subsubsection{The hierarchy of trigonometric Dynamical Models at $p=1$} The first step of reduction from the Dynamical to the Spin Models consists in taking the static limit $ p\rightarrow 1$ in the construction described in the previous subsection. Following the idea of [TH] we expand the generating function for the commuting charges of the hierarchy around the point $p=1$ up to the linear term in $p-1$: \begin{equation} \Delta (u) = \Delta_0(u) + (p-1) \Delta_1(u) + O ( (p-1)^2). \end{equation} In the symmetry generator $ T_a(u) $ we retain the leading term only: \begin{equation} T_a(u) = T_a^0(u) + O(p-1). \end{equation} The operators $ \Delta_0(u) $ and $ \Delta_1(u) $ turn out to have a very special form. First of all $ \Delta_0(u) $ is a {\em constant }: \begin{equation} \Delta_0(u) = \prod_{i=1}^N(1 + uq^{2i - N -1}). \end{equation} ({\bf Remark} In the paper [TH] which deals with the rational case, the leading term of the generating function \begin{align*} \Delta_{rational}(u) & := \prod_{i=1}^N(u - d_i) , \\ (d_i & := hD_i + \sum_{j>i}\frac{z_i}{z_i-z_j}K_{i,j}-\sum_{j<i}\frac{z_j}{z_j-z_j}K_{i,j} \quad (i=1,\dots,N) ) \end{align*} in the static limit $ h\rightarrow 0 $ is not a constant but a function of $z_1,\dots,z_N$. This is due to a choice of the Dunkl operators $d_i$ (above) which do not act in the space of polynomials. The rational limit of the affine Hecke generators $ Y_i $ which we use gives the gauge transformed Dunkl operators acting in the space of polynomials ( Cf. [BGHP]).) Since $\Delta_0(u)$ is a constant, the first-order differential operator $ \Delta_1(u) $ takes over the role of the generating function for Integrals of Motion: \begin{align} [\Delta_1(u) , \Delta_1(v) ]& = 0, \\ [\Delta_1(u) , T^0_a(v)] & = 0 . \end{align} Secondly, we find that $\Delta_1(u)$ has the following structure: \begin{equation} \Delta_1(u) = \sum_{i=1}^N \theta(u;z)_iD_i + \Xi(u;z) . \end{equation} Where $ \Xi(u;z) $ is a function of the operators $ z_1,\dots,z_N $ and $ K_{i,j} \quad (i,j = 1,\dots,N ) $ only. On the other hand the coefficients $ \theta(u;z)_i $ do not depend on the operators of coordinate permutation and are functions of the coordinates $ z_1,\dots,z_N \quad (i=1,\dots,N) $ only. This kind of separation of the differentials $ D_i $ and the operators of coordinate permutation was first observed by [TH] in the rational case. For the differential part $ {\cal D}(u) := \sum_{i=1}^N \theta(u;z)_iD_i $ of $\Delta_1(u) $ we obtain an explicit expression in terms of the generating function $ D(u;p,t)$ for the Macdonald operators [M],[JKKMP] (Cf. {\bf 2.1}): \begin{equation} {\cal D}(u) = D_1(q^{N-1}u;q^{-2}) . \end{equation} Where $ D_1(u;t)$ is the linear term in the expansion of the generating function $ D(u;p,t)$ around the point $ p=1$ : \begin{equation} D(u;p,t) = \Delta_0(u;t) + (p-1)D_1(u;t) + O((p-1)^2). \end{equation} The zero-order part $\Xi(u;z)$ of the differential operator $ \Delta_1(u) $ is the object which we use to define the hierarchy of Integrable Spin Models. First of all we observe that $ \Delta_1(u) $ lies in the centre of the Affine Hecke Algebra generated by $ g_{i,i+1} \quad (i=1,\dots,N-1) $ and $ y_i := Y_i|_{p=1} \quad (i=1,\dots,N) $. Therefore $ \Delta_1(u) $ acts in the bosonic subspace ${\cal B}$ defined in the previous subsection, and in this subspace we have analogs of the relations (0.0.17,-.19): \begin{align} [ \widehat{\Delta_1(u)} , \widehat{\Delta_1(v)}]{\cal B} & = 0 , \\ [ \widehat{\Delta_1(u)} , \widehat{T^0_a(v)} ]{\cal B} & = 0 , \\ (\bar{R}_{ab}(u/v)\widehat{T^0_a(u)}\widehat{T^0_b(v)} & - \widehat{T^0_b(v)}\widehat{T^0_a(u)}\bar{R}_{ab}(u/v)){\cal B} = 0 . \end{align} The operator $\widehat{\Delta_1(u)}$ is a sum of two parts: the $ {\cal D}(u) $ which is a first order differential operator and $ \widehat{\Xi(u;z)}$ which is a matrix acting on $H$ whose entries are rational functions of the coordinates $ z_1,\dots,z_N .$ At the second step of the reduction from the Dynamical Models to the Spin Models we shall eliminate the differential part of $\widehat{\Delta_1(u)}$ by restricting the coordinates to special values in a way which leaves the Integrability and the $ U_q'(\hat{{\frak g}{\frak l}}_2) $-invariance intact. \subsubsection{Definition of the hierarchy of trigonometric Spin Models} The way to ``freeze'' the coordinates while keeping the spins as dynamical variables goes trough the use of the {\em evaluation map } $Ev(v): {\Bbb C}[z_1,\dots,z_N]\otimes H \mapsto H $. This map is parametrized by complex numbers $ v_1,\dots, v_N $ and works by taking values of functions of $z_1,\dots,z_N $ at the point $ z_1 = v_1 ,\dots,z_N = v_N $. We use this map at the special point $z = \omega : z_1 = \omega^1,\dots,z_N = \omega^N $ where $ \omega = e^{2\pi i /N}.$ In order to explain the relevance of this point we first of all observe, that at this point the coefficients of the differential operator $ {\cal D}(u) $ are all equal one to another: \begin{equation} \theta(u;\omega)_i = \theta(u) \qquad (i=1,\dots,N) . \end{equation} Where the constant $\theta(u)$ is given by eq. (2.2.26) in the main text. Next, we observe, that in the expansion of $\widehat{\Delta_1(u)}$ in $u$: $ \widehat{\Delta_1(u)}:= \sum_{n=1}^N u^n \Hat{\Delta}_1^{(n)} $ the term $\Hat{\Delta_1^{(N)}}$ is the scale operator: $\Hat{\Delta}_1^{(N)} = D_1 + D_2 + \dots + D_N$. We can modify the generating function $ \widehat{\Delta_1(u)}$ by subtracting from it the product of the constant $\theta(u)$ and the operator $\Hat{\Delta}_1^{(N)}$. The equations (0.0.28 -.30) clearly still hold for this modified generating function $ \widehat{\Delta_1(u)} - \theta(u)\Hat{\Delta}_1^{(N)} $. Moreover due to (0.0.31) for any vector $ b \in {\cal B} $ we have: \begin{equation} Ev(\omega) (\widehat{\Delta_1(u)} - \theta(u)\Hat{\Delta}_1^{(N)})b = Ev(\omega) \widehat{\Xi(u;z)}b . \end{equation} The map $Ev(\omega)$ naturally pulls through the operators $\widehat{\Xi(u;z)}$ and $ \widehat{T_a^0(u)} $ since these operators are matrices acting in $H$, and the entries of these matrices are rational functions of $z \equiv ( z_1,\dots,z_N )$ non-singular at the point $z = \omega$. By pulling $Ev(\omega)$ through $\widehat{\Xi(u;z)}$ and $ \widehat{T_a^0(u)}$ we define operators $ \Xi(u;\omega) $ and $ T_a^0(u;\omega) $ acting in the image of $ Ev(\omega) $ in $H$ by : \begin{align} Ev(\omega)\widehat{\Xi(u;z)}{\cal B} & = \Xi(u;\omega) Ev(\omega){\cal B} , \\ Ev(\omega)\widehat{T_a^0(u)}{\cal B} & = T_a^0(u;\omega) Ev(\omega){\cal B} . \end{align} Taking (0.0.31) and (0.0.33,.34) into account we apply the evaluation map to the relations (0.0.28 - .30) and get: \begin{align} [ \Xi(u;\omega) , \Xi(v;\omega)]H_{{\cal B}}(\omega) & = 0 , \\ [ \Xi(u;\omega) , T^0_a(v;\omega) ]H_{{\cal B}}(\omega) & = 0 , \\ (\bar{R}_{ab}(u/v)T^0_a(u;\omega)T^0_b(v;\omega) & - T^0_b(v;\omega)T^0_a(u;\omega)\bar{R}_{ab}(u/v))H_{{\cal B}}(\omega) = 0 . \end{align} Where $ H_{{\cal B}}(\omega):= Ev(\omega){\cal B} \subset H .$ Next, we prove, that $H_{{\cal B}}(\omega) = H $ .Therefore the relations (0.0.35 -.37) express Integrability and $ U_q'(\hat{{\frak g}{\frak l}}_2)$-invariance of the hierarchy of Spin Chain Models. The Hamiltonians of these Models act in $H$ and are obtained by expanding the generating function $ \Xi(u;\omega) $ in the parameter $u$: \begin{equation} \Xi(u;\omega) = \sum_{n=1}^{N-1} u^n \Xi^{(n)}(\omega). \end{equation} While the generating function $ \Xi(u;\omega) $ is completely defined by the relation (0.0.25), the computation of an explicit expression is still quite a difficult task. We have computed the explicit expression only for the first member of the hierarchy -- the operator $ \Xi^{(1)}(\omega)$. To give this expression we introduce several notations. For $ i\neq j \in \{1,\dots,N-1\}$ define the rational functions: \begin{equation} a_{i,j} := \frac{q^{-1}z_i - qz_j}{z_i-z_j}\; , \quad b_{i,j} := \frac{(q-q^{-1})z_i}{z_i-z_j}. \end{equation} Define also the matrices: \begin{multline*} Y_{i,i+1}(w) := \frac{ wt_{i,i+1} - t_{i,i+1}^{-1} }{qw - q^{-1}}, \\ M^{(j,i)}(x,y) := \left( Y_{i+1,i+2}(y/z_{i+1})\dots Y_{j-1,j}(y/z_{j-1})\right)^{-1}Y_{i,i+1}(x/y) \times \\ \times \left( Y_{i+1,i+2}(x/z_{i+1})\dots Y_{j-1,j}(x/z_{j-1})\right) \quad (j > i). \end{multline*} With these notations we have: \begin{multline} \Xi^{(1)}(\omega) = \\ \sum_{M=2}^N \frac{(-1)^M}{\d} \sum_{N \geq i_M > \dots > i_1 \geq 1} {\cal H}(\omega)_{i_M,i_{M-1},\dots,i_1} R(\omega)_{i_2,i_1}R(\omega)_{i_3,i_2 \dots }R(\omega)_{i_M,i_{M-1}} + cI ; \\ {\cal H}(z)_{i_M,i_{M-1},\dots,i_1}:= \left( \prod_{i_1 < f < i_2}a_{i_1,f}\right)\left( \prod_{i_2 < f < i_3}a_{i_2,f} \right)\dots \left(\prod_{i_M < f < N + i_1}a_{i_M,f (\mod N)}\right) \times \\ \times b_{i_M,i_{M-1}} b_{i_{M-1},i_{M-2}}\dots b_{i_2,i_1} b_{i_1,i_M}, \\ R(z)_{j,i} := M^{(j,i)}(z_{i_M},z_i) \qquad (N \geq j > i \geq 1). \end{multline} Where $c$ is unimportant constant. In the limit $q \rightarrow 1 $ we recover the Haldane Shastry Hamiltonian: \begin{equation} \lim_{q\rightarrow 1}\frac{\Xi^{(1)}(\omega)}{\d} = H_{HS} := -\sum_{N \geq j > i \geq 1} \frac{\omega^i \omega^j}{ (\omega^i - \omega^j)^2 } (P_{i,j} - 1). \end{equation} While the Hamiltonian $\Xi^{(1)}(\omega)$ looks rather intimidating, its spectrum, as well as the spectrum of the whole hierarchy generated by $\Xi(u;\omega)$ has a remarkably simple additive form. We describe this spectrum in the next subsection. \subsubsection{Eigenvalue spectrum of the hierarchy of Spin Hamiltonians} As in the case of the Haldane Shastry Model the common eigenspaces of the operators $\Xi^{(n)}(\omega)\quad (n=1,\dots,N-1)$ are in one-to-one correspondence with ${\frak sl}_2$ motifs (Cf. [HHTBP] or [BPS]). A sequence of integers $(m_1,m_2,\dots,m_M) $ is called an ${\frak sl}_2$ motif iff : \begin{align} & 1 \leq m_1 < m_2 < \dots < m_M \leq N-1 \, ; \tag{\theequation a.} \\ & m_{i+1} - m_i \geq 2 \quad ( i = 1,\dots , M-1 ) \tag{\theequation b.}. \end{align} For a fixed $N$ we denote the set of all $\frak{sl}_2$ motifs including the empty one by ${\frak M}_N$. For the eigenspace of $\Xi(u;\omega)$ which correspons to a motif $(m_1,m_2,\dots,m_M) $ we use the notation $ H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega)$. We have: \begin{equation} \Xi(u;\omega) H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega) = \left( \sum_{i=1}^M \xi^{(m_i)}(u) \right)H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega).\end{equation} Where the sum is understood to be zero for $ M=0$. \\ The elementary eigenvalue $\xi^{(m)}(u) \quad (m \in \{1,\dots,N-1\})$ is \begin{equation} \xi^{(m)}(u) = u\prod_{k=1}^N(1 + u q^{2k-N-1})\left\{\sum_{i=1}^m \frac{q^{2i - N-1}}{1+ uq^{2i - N-1}} -\frac{m}{N} \sum_{i=1}^N \frac{q^{2i - N-1}}{1+ uq^{2i - N-1}} \right\}. \end{equation} In particular the elementary eigenvalue of the Hamiltonian $\Xi^{(1)}(\omega)$ is given by \begin{equation} \xi^{(m),(1)} = \frac{q^{-N}}{N}( Nq^m[m]_q - mq^N[N]_q ). \end{equation} Where we used the usual notation $ [x]_q \equiv \frac{q^x - q^{-x}}{\d}$. In the limit $ q\rightarrow 1 $ we recover the elementary eigenvalue of the Haldane Shastry Model [HHTBP,BPS]: \begin{equation} \lim_{q\rightarrow 1}\frac{\xi^{(m),(1)}}{\d} = m(m - N). \end{equation} The space of states of the Spin Models is represented as a direct sum of the eigenspaces $ H^{(m_1,m_2,\dots,m_M)}(\omega)$: \begin{equation} H = \bigoplus_{(m_1,m_2,\dots,m_M) \in {\frak M}_N }H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega). \end{equation} \subsubsection{Structure of the common eigenspaces of the hierarchy of Spin Hamiltonians.} Each of the eigenspaces $ H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega)$ is an irreducible highest-weight representation of $U_q'(\hat{{\frak g}{\frak l}}_2)$. The Drinfeld polynomial [CP] $Q^{(m_1,m_2,\dots,m_M)}(u)$ of $ H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega)$ is \begin{equation} Q^{(m_1,m_2,\dots,m_M)}(u) = \prod\begin{Sb}1\leq k \leq N \\ k \not\in \{m_i,m_i+1\} \end{Sb} (1 - q^{-2k+N+1}u). \end{equation} (Cf. {\bf 1.2 } for our conventions about Drinfeld polynomials ). To describe an eigenspace $ H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega) \quad (m_1,m_2,\dots,m_M) \in {\frak M}_N $ in a more explicit way we give some facts (Cf. {\bf 4.1}) about eigenvectors of the operator $ \Delta_1(u) $ which defines the hierarchy of Dynamical Models at $p=1$. The linear space of polynomials ${\Bbb C}[z_1,\dots,z_N]$ is represented as a direct sum of eigenspaces of the operator $\Delta_1(u) $. There is a one-to-one correspondence between these eigenspaces and partitions with $ N$ parts. In an eigenspace which corresponds to a partition $ \l := ( \l_1 \geq \l_2 \geq \dots \l_N \geq 0 ) $ there exists a basis $\{ \varphi_{\sigma}^{\l}(z) \}_{\sigma \in \S}$. The polynomials forming this basis are indexed by elements from symmetric group $S_N$. There is a one-to-one correspondence between the elements of the basis $\{ \varphi_{\sigma}^{\l}(z) \}_{\sigma \in \S}$ and elements of the set $ \S \quad (\S \subset S_N )$ defined in {\bf 1.4}. A polynomial $ \varphi_{\sigma}^{\l}(z) \quad (\sigma \in \S) $ is completely specified by the three conditions:\begin{align} \Delta_1(u) \varphi_{\sigma}^{\l}(z) & = \left(\prod_{k=1}^N(1 + u q^{2k-N-1})\left\{\sum_{i=1}^N \frac{uq^{2i-N-1}}{1 + uq^{2i-N-1}}\l_i \right\} \right) \varphi_{\sigma}^{\l}(z) , \tag{\theequation a.} \\ y_i \varphi_{\sigma}^{\l}(z) & = q^{2\sigma_i - N -1} \varphi_{\sigma}^{\l}(z) \quad (i=1,\dots,N) , \tag{\theequation b.} \\ \varphi_{\sigma}^{\l}(z) & = z_1^{\l_{\sigma_1}}z_2^{\l_{\sigma_2}} \dots z_N^{\l_{\sigma_N}} + \text{smaller monomials}.\tag{\theequation c.} \end{align} We remind, that $y_i := Y_i|_{p=1} \quad (i=1,\dots,N) $. \\ The ``smaller monomials `` means a linear combination of monomials that are smaller than the monomial $z_1^{\l_{\sigma_1}}z_2^{\l_{\sigma_2}} \dots z_N^{\l_{\sigma_N}}$ in the ordering described in {\bf 1.4(2)} (Cf. [BGHP]). With any motif $ (m_1,m_2,\dots,m_M) \in {\frak M}_N $ associate the partition \begin{multline} (\underset{1}{M},\dots,\underset{m_1}{M},\underset{m_1+1}{M-1},\dots ,\underset{m_2}{M-1},\underset{m_2+1}{M-2}, \ldots \\ \ldots, \underset{m_M}{1},\underset{m_M+1}{0},\dots ,\underset{N}{0} ) \end{multline} We use the same notation $ (m_1,m_2,\dots,m_M) $ for a motif and the associated partition. For an element $\sigma $ of the symmetric group $S_N$ define \begin{align} \{\sigma_1 , \sigma_2 , \dots ,\sigma_N \} & := \sigma.\{1,2,\dots,N \} \, ; \tag{\theequation a.} \\ i & := \sigma_{p^{\sigma}_i} \qquad (i=1,\dots,N) . \tag{\theequation b.} \end{align} For any $(m_1,m_2,\dots,m_M) \in {\frak M}_N $ define the subset $S^{(m_1,m_2,\dots,m_M)}_{N,(m_1,m_2,\dots,m_M)}$ of $S_N$ (Cf. {\bf 4.3(2)}): \begin{multline} S^{(m_1,m_2,\dots,m_M)}_{N,(m_1,m_2,\dots,m_M)} := \\ \left\{ \sigma \in S_N \quad \begin{array}{| c c c } p^{\sigma}_i < p^{\sigma}_{i+1} & ( m_k < i < m_{k+1}) & \text{for all}\; k \in \{0,1,\dots,M\} \\ p^{\sigma}_{m_k} > p^{\sigma}_{m_k+1} & & \text{for all} \; k \in \{1,\dots,M-1\} \end{array} \right\}. \end{multline} Where we adopt the convention: $ m_0 := 0 \, , m_{M+1} := N+1 .$ Next, for any motif $ (m_1,m_2,\dots,m_M) \in {\frak M}_N $ we define the subspace $W^{(m_1,m_2,\dots,m_M)}$ of the space of states $H$ (Cf. {\bf 6.1(2) }) as follows: \begin{multline} W^{(m_1,m_2,\dots,m_M)}:= \\ S_q(\underset{1}{V}\otimes\dots\otimes\underset{m_1-1}{V})\otimes A_q(\underset{m_1}{V}\otimes\underset{m_1+1}{V})\otimes S_q(\underset{m_1+2}{V}\otimes\dots\otimes\underset{m_2-1}{V})\otimes A_q(\underset{m_2}{V}\otimes\underset{m_2+1}{V})\otimes \ldots \\ \ldots \otimes S_q(\underset{m_{M-1}+2}{V}\otimes\dots\otimes\underset{m_M-1}{V})\otimes A_q(\underset{m_M}{V}\otimes\underset{m_M+1}{V})\otimes S_q(\underset{m_M+2}{V}\otimes\dots\otimes\underset{N}{V}) \\ W^{(m_1,m_2,\dots,m_M)} \subset H:= \underset{1}{V}\otimes\underset{2}{V}\otimes \ldots \underset{N}{V}. \end{multline} Where $ S_q $ and $ A_q$ mean $q$-symmetrization and $q$-antisymmetrization as defined in (1.1.14,.15,5.5.17). The space $W^{(m_1,m_2,\dots,m_M)} \quad (m_1,m_2,\dots,m_M) \in {\frak M}_N$ is an irreducible highest weight $U_q'(\hat{{\frak g}{\frak l}}_2)$-module. The Drinfeld polynomial of this module is given by (0.0.47). The $U_q'(\hat{{\frak g}{\frak l}}_2)$-action on $W^{(m_1,m_2,\dots,m_M)}$ is given by \begin{equation} {\cal L}(u;\{q^{2\sigma[0]_i - N -1}\}):= L_{a1}(uq^{2\sigma[0]_1 - N -1})L_{a2}(uq^{2\sigma[0]_2 - N -1})\dots L_{aN}(uq^{2\sigma[0]_N - N -1}). \end{equation} Where for a fixed $(m_1,m_2,\dots,m_M)$ we have introduced the notation \begin{equation} \sigma[0] := (m_1,m_1+1)(m_2,m_2+1)\dots(m_M,m_M+1) \in S_{N,(m_1,m_2,\dots,m_M)}^{(m_1,m_2,\dots,m_M)} \subset S_N. \end{equation} The eigenspace $ H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega) \quad ((m_1,m_2,\dots,m_M) \in {\frak M}_N) $ is isomorphic to $ W^{(m_1,m_2,\dots,m_M)}$ as $U_q'(\hat{{\frak g}{\frak l}}_2)$-module. This isomorphism is given by an invertible intertwiner $ \check{\BU }^{(m_1,m_2,\dots,m_M)}(\omega)$ \begin{equation} H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega) = \check{\BU }^{(m_1,m_2,\dots,m_M)}(\omega)W^{(m_1,m_2,\dots,m_M)}. \end{equation} The intertwiner $\check{\BU }^{(m_1,m_2,\dots,m_M)}(\omega)$ is defined by the expression \begin{equation} \check{\BU}^{(m_1,m_2,\dots,m_M)}(\omega) := (-(q^2 + 1))^M\sum_{\sigma \in S_{N,(m_1,m_2,\dots,m_M)}^{(m_1,m_2,\dots,m_M)}}\varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}(\omega){\Bbb Y}(\sigma) . \end{equation} In this definition \begin{equation} \varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}(\omega) = \varphi_{\sigma}^{\l}(z)|_{z_1=\omega^1,\dots,z_N=\omega^N}. \end{equation} Where the partition $ \l$ is the one specified by $ (m_1,m_2,\dots,m_M) $ in accordance with (0.0.48). The matrix ${\Bbb Y}(\sigma) \quad ( \sigma \in S_{N,(m_1,m_2,\dots,m_M)}^{(m_1,m_2,\dots,m_M)})$ is an intertwiner which is defined by the recursion relations (Cf. {\bf 5.2(4)} ): \begin{gather*} {\Bbb Y}(\sigma[0]):= \text{Id} \; , \\ {\Bbb Y}((i,i+1)\sigma) := \begin{cases} Y^+_{i,i+1}(q^{2\sigma_i-2\sigma_{i+1}}){\Bbb Y}(\sigma) & \text{ if $ \sigma_i - \sigma_{i+1} \geq 2 $}, \\ Y^-_{i,i+1}(q^{2\sigma_i-2\sigma_{i+1}}){\Bbb Y}(\sigma) & \text{ if $ \sigma_i - \sigma_{i+1} \leq -2 $ }. \end{cases} \end{gather*} Where the matrices $ Y^{\pm}_{i,i+1}(w) $ are \begin{align} Y^{\pm}_{i,i+1}(w) & := \varrho^{\pm}(w)\frac{wt_{i,i+1} - t_{i,i+1}^{-1}}{q^{-1}w - q } \, , \\ \varrho^+(w) & := \frac{w -1}{q^2 w -1} \;, \qquad \varrho^-(w) := \frac{w - q^2}{ w -1}. \notag \end{align} Notice that ${\Bbb Y}(\sigma) \quad ( \sigma \in S_{N,(m_1,m_2,\dots,m_M)}^{(m_1,m_2,\dots,m_M)})$ is an invertible intertwiner. In general we cannot claim to know the explicit expression for the intertwiner $\check{\BU}^{(m_1,m_2,\dots,m_M)}(\omega)$ since we have not found the eigenvectors $ \varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}(z) $ explicitely. One exception is the case $q=0$ when $\check{\BU}^{(m_1,m_2,\dots,m_M)}(\omega)$ becomes very simple. In this case we have \begin{equation} \check{\BU}^{(m_1,m_2,\dots,m_M),q=0}(\omega)|_{W^{(m_1,m_2,\dots,m_M),q=0}} = \omega^{\frac{1}{2}\sum_{i=1}^Mm_i(m_i + 1)}{\mathrm {Id}}. \end{equation} Therefore at $q=0$ the eigenspace $H_{{\cal B}}^{(m_1,m_2,\dots,m_M)}(\omega)$ is the linear span of the following vectors in $H$: \begin{multline*} \left|\left\{\begin{array}{c c c c c }\overset{1}{+}&\overset{}{+}&\overset{}{+}&\dots &\overset{m_1-1}{+} \\ -& +& +& \dots & + \\ -& -& +& \dots & + \\ \vdots & \vdots & \vdots & & \vdots \\ - & - & - & \dots & - \end{array} \right\} \right. \overset{m_1}{+}\overset{m_1+1}{-} \left\{\begin{array}{c c c c c }\overset{m_1+2}{+}&\overset{}{+}&\overset{}{+}&\dots &\overset{m_2-1}{+} \\ -& +& +& \dots & + \\ -& -& +& \dots & + \\ \vdots & \vdots & \vdots & & \vdots \\ - & - & - & \dots & - \end{array} \right\} \overset{m_2}{+}\overset{m_2+1}{-} \ldots \\ \ldots \overset{m_M}{+}\overset{m_M+1}{-} \left.\left\{\begin{array}{c c c c c }\overset{m_M+2}{+}&\overset{}{+}&\overset{}{+}&\dots &\overset{N}{+} \\ -& +& +& \dots & + \\ -& -& +& \dots & + \\ \vdots & \vdots & \vdots & & \vdots \\ - & - & - & \dots & - \end{array} \right\} \right\rangle \end{multline*} Where we use the notation: \begin{equation} |\epsilon_1 \epsilon_2 \dots \epsilon_N \rangle := v^{\epsilon_1}\otimes v^{\epsilon_2}\otimes \dots \otimes v^{\epsilon_N} \quad ( \epsilon_i = \pm ) {}. \end{equation} For example when $N$ is even, $H_{{\cal B}}^{(1,3,\dots,N-1)}(\omega)$ is one-dimensional and at $q=0$ it is spanned by the vector (antiferromagnetic ground state): \begin{equation*} |+ - + - \dots + - \rangle \end{equation*} \mbox{} \noindent In the rest of the paper we give a detailed exposition of the matters which were briefly recounted in this introduction. In the sec.1 we gather predominantly known facts about the trigonometric Dynamical Models of [BGHP] and explain the conventions about the algebra $ U_q'(\hat{{\frak g}{\frak l}}_2) $ that we use. In the sec. 2 we discuss the hierarchy of Dynamical Models in the static limit $ p=1$. In the sec. 3 we define the hierarchy of the Spin Models. The sec. 4. is concerned with properties of the eigenvectors of the hierarchy of Dynamical Models in the limit $p = 1$. In the sec. 5 we construct the Hecke-invariant ``bosonic'' eigenspaces for the Dynamical Models with spin at $p=1$. The eigenvalue spectrum of the Spin Models and the $U_q'(\hat{{\frak g}{\frak l}}_2) $-representation content of their eigenspaces are derived in sec. 6. \mbox{} \noindent \begin{Large}{\bf Acknowledgments}\end{Large}\\ I am most grateful to Drs. R. Kedem and R. Weston and to Professors M.Jimbo, M.Kashiwara and T.Miwa for numerous discussions and support. \section{The $U_q'(\hat{{\frak g}{\frak l}}_2)$-invariant Dynamical Models} In this section we summarize largely known facts about the trigonometric Dynamical Models defined by [BGHP]. We also recount several facts about the algebra $U_q'(\hat{{\frak g}{\frak l}}_2)$ and explain our notations. \subsection{The representations of the Affine Hecke Algebra} \subsubsection{The representation of $ \widehat{H_N(q)} $ in the space of polynomials} Following [BGHP] define the operators $ g_{i,j} \in End( {\Bbb C}[z_1,\dots,z_N]) \quad (i,j =1,\dots, N) $: $$ g_{i,j} := a_{i,j}K_{i,j}+ b_{i,j} $$ where $$ a_{i,j} = \frac{q^{-1}z_i - qz_j}{z_i-z_j}\;, \qquad b_{i,j} = \frac{(q-q^{-1})z_i}{z_i-z_j} = q - a_{i,j}. $$ $ K_{i,j} $ is the interchange operator for variables $ z_i , z_j $. For $ p \in {\Bbb C} $ and $ D_i := z_i \frac{\partial}{\partial z_i} $ define operators $ Y_i \in End( {\Bbb C}[z_1,\dots,z_N]) (i=1,\dots,N) $: \begin{eqnarray} Y_i & := & g_{i,i+1}^{-1}K_{i,i+1}\dots g_{i,N}^{-1}K_{i,N}p^{D_i}K_{1,i}g_{1,i}\dots K_{i-1,i}g_{i-1,i} \end{eqnarray} Taken together with $ { g_{i,i+1} }\; (i=1,\dots, N-1 )$ these operators satisfy the relations of the Affine Hecke Algebra $ \widehat{H_N(q)} $ [BGHP] : \begin{eqnarray} g_{i,i+1}^2 & = & (q-q^{-1})g_{i,i+1} + 1 \\ g_{i,i+1} g_{k,k+1} & = & g_{k,k+1} g_{i,i+1} \; , \qquad |i-k| \geq 2 \\ g_{i,i+1} g_{i+1,i+2}g_{i,i+1} & = & g_{i+1,i+2}g_{i,i+1} g_{i+1,i+2} \\ Y_k g_{i,i+1} & = & g_{i,i+1} Y_k \; , \qquad k \neq i,i+1 \\ g_{i,i+1} Y_i & = & Y_{i+1} g_{i,i+1}^{-1} \\ Y_i Y_j & = & Y_j Y_i \: . \end{eqnarray} Any symmetric polynomial in $ Y_i \; (i=1,\dots,N) $ belongs to the center of $ \widehat{H_N(q)} $. All symmetric polynomials in $ Y_i \; (i=1,\dots,N) $ are generated by the elementary symmetric polynomials which are obtained by expanding in the parameter $u$ the generating function $ \Delta(u) $ : \begin{eqnarray} \Delta(u) & = & \prod_{(i=1,\dots,N) }(1 + uY_i) \end{eqnarray} \subsubsection{The representation of the Hecke Algebra $ H_N(q) $ in $ ({\Bbb C}^2)^{\otimes N} $ } Let $ V := {\Bbb C}^2 = {\rm span} \{ v^+,v^- \} $. Define $ t \in End(V\otimes V) $ by \begin{eqnarray} t v^+\otimes v^- & = & (q-q^{-1})v^+\otimes v^- + v^-\otimes v^+ \\ t v^-\otimes v^+ & = & v^+\otimes v^- \\ t v^\pm\otimes v^\pm & = & qv^\pm\otimes v^\pm \end{eqnarray} The operators $ \Pi^\pm(q) $: \begin{eqnarray} \Pi^+(q) & := & \frac{ q^{-1} + t }{q + q^{-1} } \\ \Pi^-(q) & := & \frac{ q - t }{q + q^{-1} } \end{eqnarray} are orthogonal projectors on the subspaces: \begin{eqnarray} S_q( V\otimes V ) & := & {\Bbb C} \{ v^+\otimes v^+ , v^-\otimes v^- , qv^+\otimes v^- + v^-\otimes v^+ \} \\ A_q( V\otimes V ) & := & {\Bbb C} \{ v^+\otimes v^- - q v^-\otimes v^+ \} \end{eqnarray} respectively. Let $ H := V^{\otimes N}$. For an $ O \in End(V\otimes V) $ denote by $ O_{i,j} \in End(H) $ the standard injection $ End(V\otimes V) \rightarrow End(H) $ which acts trivially on all the factors except the $i$-th and $j$-th ones. The matrices $ t_{i,i+1} \quad (i=1,\dots, N-1) $ satisfy the defining relations (1.1.2-.4) of the finite-dimensional Hecke Algebra $ H_N(q) $. \subsection{The Algebra $ U_q'(\hat{{\frak g}{\frak l}}_2) $ at level 0 and some of its representations} In this subsection we summarize several facts about the algebra $U_q'(\hat{{\frak g}{\frak l}}_2)$ and its representations. Our conventions and notations mainly follow [JKKMP]. \subsubsection{The Algebra $ U_q'(\hat{{\frak g}{\frak l}}_2) $ } In the $L$-operator formalism $ U \equiv U_q'(\hat{{\frak g}{\frak l}}_2) $ at zero level is defined to be the associative algebra with unit generated by elements $ l_{ij}^\pm[\pm n]\;(i,j = 1,2 \:; n = 0,1,\dots) $.The $L$-operators $ L^\pm(u) \in End(V\otimes U) $ are the generating series in the spectral parameter $u$: \begin{eqnarray} L^\pm(u) & := & \sum_{\pm n \geq 0 } u^n \left( \begin{array}{c c} l_{11}[n] & l_{12}[n] \\ l_{21}[n] & l_{22}[n] \end{array} \right) \end{eqnarray} The defining relations of $U$ are written in the the form: \begin{eqnarray} \bar{R}_{ab}(u/v)L^\pm_a(u)L^\pm_b(v) & = & L^\pm_b(v)L^\pm_a(u)\bar{R}_{ab}(u/v) \\ \bar{R}_{ab}(u/v)L^+_a(u)L^-_b(v) & = & L^-_b(v)L^+_a(u)\bar{R}_{ab}(u/v) \\ l_{ii}^+[0]l_{ii}^-[0]\; = \; 1 \quad (i=1,2)\:, & & l_{21}^+[0] \: = \: l_{12}^-[0] \: = \: 0 \end{eqnarray} where the $R$-matrix $ \bar{R}_{ab}(z) \in End(V\otimes V := V_a \otimes V_b ) $ is defined as follows: \begin{eqnarray} \bar{R}(z) & = & \frac{ zt - t^{-1} }{qz - q^{-1}}P \end{eqnarray} by $P$ we denote the permutation operator in $ V\otimes V $. \subsubsection{Some representations of $U$} A finite-dimensional highest weight module $W$ of $U$ contains non-zero vector $\Omega$ which satisfies the condition: \begin{eqnarray} L^\pm(u)\Omega & = & \left( \begin{array}{c c} A^\pm(u) & * \\ 0 & D^\pm(u) \end{array} \right) \Omega \end{eqnarray} where $ A^\pm(u)$ and $ D^\pm(u)$ are ${\Bbb C}$-valued series in $u$. If $W$ is irreducible, it is specified up to equivalence by its Drinfeld polynomial $Q(u)$ whch is determined by the conditions: $ Q(0) \: = \: 1 $ and \begin{eqnarray} q^{degQ}\frac{Q(q^{-2}u)}{Q(u)} & = & \frac{A^+(u^{-1})}{D^+(u^{-1})} \qquad ( u \rightarrow 0 ) \\ & = & \frac{A^-(u^{-1})}{D^-(u^{-1})} \qquad ( u \rightarrow \infty ) \end{eqnarray} The example of such $W$ is the 2-dimensional evaluation module $ W(a) $ where $ a \in {\Bbb C} \backslash \{0\} $ is the parameter. As a vector space $W(a)$ is isomorphic to $V$. The generators $ l_{ij}^\pm[\pm n]\;(i,j = 1,2 ; n = 0,1,\dots) $ are defined by expanding the $L$-operator: \begin{eqnarray} L(ua) & := & \frac{uat - t^{-1}}{ua - 1}P \; \in End(V_a \otimes V) \end{eqnarray} into power series in $ u^\pm $ around zero and infinity respectively. The Drinfeld polynomial of $W(a)$ is: $ Q(u;a) \; = \; 1 - a^{-1}u .$ A pair of tensor products $W(a)\otimes W(b)$, $W(b)\otimes W(a)$ is intertwined by the matrix \begin{eqnarray} \bar{Y}(z) & = & zt - t^{-1} \quad ( z \in {\Bbb C} ) \; \in End(V\otimes V ) \end{eqnarray} i.e.: \begin{eqnarray} \bar{Y}_{12}(a/b)L_{a1}(ua)L_{a2}(ub) & = & L_{a1}(ub)L_{a2}(ua)\bar{Y}_{12}(a/b) \end{eqnarray} this relation holds in the tensor product of an auxiliary copy of $V$ indicated by the subscript $a$ and $ V \otimes V $ indicated by subscripts 1 and 2. The intertwiner $\bar{Y}(a/b)$ is invertible unless either $ a = q^2b $, in which case \begin{eqnarray} \bar{Y}(q^2) & = & (q^2 - 1)(q+q^{-1}) \Pi^+(q) \end{eqnarray} or $ a = q^{-2}b $, in which case \begin{eqnarray} \bar{Y}(q^{-2}) & = & (1-q^{-2})(q+q^{-1}) \Pi^-(q) \end{eqnarray} Together with (26) this leads to the invariance relations: \begin{eqnarray} L_{a1}(q^2u)L_{a2}(u):A_q(V\otimes V) & \subset & A_q(V\otimes V) \\ L_{a1}(q^2u)L_{a2}(u):S_q(V\otimes V) & \subset & A_q(V\otimes V)\oplus S_q(V\otimes V)\\ L_{a1}(u)L_{a2}(q^2u):S_q(V\otimes V) & \subset & S_q(V\otimes V) \\ L_{a1}(u)L_{a2}(q^2u):A_q(V\otimes V) & \subset & A_q(V\otimes V)\oplus S_q(V\otimes V) \end{eqnarray} \subsection{The hierarchy of $U$-invariant Dynamical Models} The central elements of $\widehat{H_N(q)}$ generated by $\Delta(u)$ were proposed in [BGHP] to define the hierarchy of integrable Dynamical Models which are trigonometric - that is $U$-invariant - generalizations of the Yangian-invariant Dynamical Models found by the same authors. Define $ T_a(u) \in End({\Bbb C}[z_1,\dots,z_N]\otimes H) $ by taking the tensor product of the $L$-operators (24): \begin{eqnarray} T_a(u)&=& L_{a1}(uY_1)L_{a2}(uY_2)\dots L_{aN}(uY_N) \end{eqnarray} After expansion in $u^\pm$, $ T_a(u) $ gives rise to a representation of $U$ in $ {\cal P} \: := \: {\Bbb C}[z_1,\dots,z_N] \otimes H $. The action of $\Delta(u)$ naturally extends to $ {\cal P}$. Retain the same notation $\Delta(u)$ for this extension. The $U$-invariance and integrability relations for $\Delta(u)$ are immediate: \begin{eqnarray} [ \Delta(u) , T_a(v) ] & = & 0 \\ \mbox{} [ \Delta(u) , \Delta(v) ] & = & 0 \end{eqnarray} Both $\Delta(u)$ and $ T_a(v)$ act in the (``bosonic'') subspace ${\cal B}$: \begin{eqnarray} {\cal B} & := & \{ b \in {\cal P} | g_{i,i+1} b = t_{i,i+1}b \; (i=1,\dots,N-1) \} \end{eqnarray} This allows to restrict $\Delta(u)$ and $ T_a(v)$ on ${\cal B}$ where both of these operators can be rewritten in such a way that they do not depend explicitely on the operators $K_{i,j}$. \subsection{Eigenvalue spectrum of the operators $ Y_i \; (i=1,\dots,N) $ } {\bf 1.} We shall work in the monomial basis of $ {\Bbb C}[z_1,\dots,z_N] $. Introduce a convenient parametrization of the elements of this basis. With a monomial $ z^{\nu} := z_1^{\nu_1}z_2^{\nu_2}\dots z_N^{\nu_N} \quad (\nu_i \in {\Bbb Z}^+ (i=1,\dots,N) ) $ associate a partition $ \l := (\l_1 \geq \l_2 \geq \dots \geq \l_N \geq 0 ) $ such that $ \{\nu_1,\nu_2,\dots,\nu_N \} = \{ \l_{\sigma_1},\l_{\sigma_2},\dots,\l_{\sigma_N} \}:= \sigma.\{ \l_1,\l_2,\dots,\l_N \} $ for some $ \sigma \in S_N = $ symmetric group with $ N-1 $ generators. A partition $ \l $ is uniquely specified by $ \nu $, while in general $ \sigma $ is not. For $ \sigma \in S_N : \{ \sigma_1,\sigma_2,\dots , \sigma_N \} = \sigma.\{1,2,\dots,N \} $ define $ p^{\sigma}_i \; : \; i = \sigma_{p^{\sigma}_i} \quad (i=1,\dots,N) .$ Let $ \Lambda_N $ be the set of all {\it N-}member partitions. For any $\l \in \Lambda_N $ write: $\l = (\l_1=\l_2=\dots =\l_{m_1} > \l_{m_1+1}=\l_{m_1+2}=\dots =\l_{m_2} > \ldots \\ \ldots > \l_{m_M+1}=\l_{m_M+2}=\dots=\l_N )$. \begin{df} For any $ \l \in \Lambda_N $: \begin{align*} \S := \{ \sigma \in S_N \; | \; p^{\sigma}_1 < p^{\sigma}_2 < \dots p^{\sigma}_{m_1} , \; p^{\sigma}_{m_1+1} < p^{\sigma}_{m_1+2} < \dots p^{\sigma}_{m_2}, \ldots \\ \ldots , p^{\sigma}_{m_M+1} < p^{\sigma}_{m_M+2} < \dots p^{\sigma}_N \}. \end{align*} \end{df} For a given $\l$ the elements of the set $ \S $ are in one-to-one correspondence with distinct rearrangements of the sequence $ \{ \l_1,\l_2,\dots,\l_N \}.$ If $ \sigma \in \S $ then $ (i\,i+1)\sigma \in \S $ iff $ \l_{\sigma_i} \neq \l_{\sigma_{i+1}} \quad (i=1,\dots,N-1) .$ The particular choice of $ \S $ as a subset of $ S_N $ parametrizing distinct rearrangements of a partition will be explained by the Proposition 1. Denote $ z^{\l_{\sigma}} := z_1^{\l_{\sigma_1}}z_2^{\l_{\sigma_2}}\dots z_N^{\l_{\sigma_N}} \quad (\l \in \Lambda_N\:, \; \sigma \in \S ).$ With this notation: \begin{eqnarray} {\Bbb C}[z_1,\dots,z_N] = \bigoplus_{ \l \in \Lambda_N } \oplus_{ \sigma \in \S } {\Bbb C} z^{\l_{\sigma}}. \end{eqnarray} {\bf 2.} Introduce an ordering on the set of monomials $ \{ z^{\l_{\sigma}} \} \quad ( \l \in \Lambda_N,\; \sigma \in \S ) .$ We say that $ \l > \tilde{\l} \quad ( \l,\tilde{\l} \in \Lambda_N ) $ iff the first ( counting from left ) non-vanishing element of the sequence $ \{ \l_1 - \tilde{\l}_1, \l_2 - \tilde{\l}_2, \dots, \l_N - \tilde{\l}_N \} $ is positive. Fix $ \l \in \Lambda_N .$ We say that $ \sigma > \tilde{\sigma}\quad ( \sigma, \tilde{\sigma} \in \S )$ iff the last non-vanishing element of the sequence $ \{ \l_{\sigma_1} - \l_{\tilde{\sigma}_1}, \l_{\sigma_2} - \l_{\tilde{\sigma}_2}, \dots \l_{\sigma_N} - \l_{\tilde{\sigma}_N} \}$ is negative. For $ \l , \tilde{\l} \in \Lambda_N \:;\; \sigma \in \S , \: \tilde{\sigma} \in S^{\tilde{\l}}_N $ define $ \l_{\sigma} > \tilde{\l}_{\tilde{\sigma}} $ iff either $ \l > \tilde{\l} \:,$ or $ \l = \tilde{\l}\:, \sigma > \tilde{\sigma} $. The ordering on monomials $ z^{\l_{\sigma}} $ is induced by the ordering on the exponents $ \l_{\sigma} \quad ( \l \in \Lambda_N,\; \sigma \in \S ) .$ \mbox{} \noindent {\bf 3.} The action of $ H_N(q) $ in the monomial basis is found by a straightforward computation to be as follows: \begin{equation} g_{i,j} z^{\l_{\sigma}} = ( i < j ) = \begin{cases} (q-q^{-1}) z^{\l_{\sigma}} + qz^{\l_{(ij)\sigma}} + \text{``s.p.''} & \text{ if $\l_{\sigma_i} > \l_{\sigma_j}$ }, \\ qz^{\l_{\sigma}} & \text{ if $\l_{\sigma_i} = \l_{\sigma_j}$ }, \\ q^{-1}z^{\l_{(ij)\sigma}} + \text{``s.p.''} & \text{ if $\l_{\sigma_i} < \l_{\sigma_j}$ }. \end{cases} \end{equation} And \begin{equation} g_{i,j}^{-1} z^{\l_{\sigma}} = (i < j ) = \begin{cases} qz^{\l_{(ij)\sigma}} + \text{``s.p.''} & \text{ if $\l_{\sigma_i} > \l_{\sigma_j}$ },\\ q^{-1}z^{\l_{\sigma}} & \text{ if $\l_{\sigma_i} = \l_{\sigma_j}$ }, \\ (q-q^{-1}) z^{\l_{\sigma}} + q^{-1}z^{\l_{(ij)\sigma}} + \text{``s.p.''} & \text{ if $\l_{\sigma_i} < \l_{\sigma_j}$ }. \end{cases} \end{equation} Where ``s.p.'' means a linear combination of monomials with smaller partitions. It follows that: \begin{equation} K_{i,j}g_{i,j} z^{\l_{\sigma}} = ( i < j ) = \begin{cases} qz^{\l_{\sigma}} + \text{``s.m.''} & \text{ if $\l_{\sigma_i} \geq \l_{\sigma_j}$ }, \\ q^{-1}z^{\l_{\sigma}} + \text{``s.m.''} & \text{ if $\l_{\sigma_i} < \l_{\sigma_j}$ }. \end{cases} \end{equation} And \begin{equation} g_{i,j}^{-1}K_{i,j} z^{\l_{\sigma}} = ( i < j ) = \begin{cases} q^{-1}z^{\l_{\sigma}} + \text{``s.m.''} & \text{ if $\l_{\sigma_i} \geq \l_{\sigma_j}$ }, \\ qz^{\l_{\sigma}} + \text{``s.m.''} & \text{ if $\l_{\sigma_i} < \l_{\sigma_j}$ }. \end{cases} \end{equation} Where ``s.m.'' signifies a linear combination of smaller monomials. \mbox{} \noindent {\bf 4.} The formulas of the preceding paragraph lead to the following proposition: \begin{prop} The operators $ Y_i \; (i=1,\dots,N) $ are triangular in the monomial basis of $ {\Bbb C}[z_1,\dots,z_N] $. The action of these operators on monomials is given by: \begin{equation*} Y_iz^{\l_{\sigma}} = p^{\l_{\sigma_i}}q^{l_{\sigma_i}}z^{\l_{\sigma}} + \text{{\em``s.m.''}} \quad ( \l \in \Lambda_N,\; \sigma \in \S,\; (i=1,\dots,N) ) \end{equation*} where $ l_i := 2i-N-1 \quad (i=1,\dots,N) $. \end{prop} This proposition shows that $ \zeta_i^{\l}(\sigma) := p^{\l_{\sigma_i}}q^{l_{\sigma_i}} \quad (\l \in \Lambda_N,\; \sigma \in \S) $ constitute a complete set of characteristic numbers of the operator $ Y_i \quad (i \in \{ 1,2,\dots,N \}).$ In order to prove that the operators $ Y_i \; (i=1,\dots,N) $ are simultaneously diagonalizable and that $ \{\zeta_i^{\l}(\sigma)\}\quad (\l \in \Lambda_N,\; \sigma \in \S) $ form the complete set of eigenvalues of $ Y_i \quad (i \in \{ 1,2,\dots,N \}) $ we shall make use of Lemma 1 discussed in the next paragraph. \mbox{} \noindent {\bf 5.} The aim of this paragraph is to recall the following ( presumably well-known ) result: \begin{lemma} Let $ {\cal V} = {\Bbb C}\{f_a\}_{a = 1,2,\dots,d={\mathrm {dim}}{\cal V}} $ be a finite-dimensional vector space. Let $ {\cal Z}_i \in End({\cal V})\quad (i=1,2,\dots , {\cal N}) $ and : \begin{align} [{\cal Z}_i,{\cal Z}_j] & = 0 \quad (i,j = 1,2,\dots, {\cal N}), \tag{a} \\ {\cal Z}_i \quad (i=1,2,\dots , {\cal N}) \quad & \text{are simultaneously triangular in the basis}\: \{f_a\}_{a = 1,2,\dots,d}: \notag \\ {\cal Z}_i f_a & = \xi_i^a f_a + \sum_{b < a} m_i^{b\,a} f_b \quad (i=1,2,\dots , {\cal N}), \tag{b} \\ & \text{where $ m_i^{b\,a} $ are coefficients.} \notag \end{align} {\em (c)} The joint set of characteristic numbers $ \{ \xi_i^a \} \quad (i=1,2,\dots,{\cal N} \:;\; a = 1,2,\dots,d ) $ is multiplicity-free: \begin{align} \forall\; a \neq b \; ( a,b = 1,2,\dots,d )\; & \; \exists \; I(a,b) \subset \{1,2,\dots,{\cal N}\}: \notag \\ \forall \; i \in I(a,b) \; & \xi_i^a - \xi_i^b \neq 0 . \notag \end{align} \noindent Then $ \exists $ a basis $ \{ \phi_a \}_{a=1,2,\dots,d} $ : \begin{align} {\cal Z}_i \phi_a & = \xi_i^a \phi_a \quad ( a=1,2,\dots,d ;\: i=1,2,\dots,{\cal N} ) \notag \\ \phi_a & = f_a + \sum_{b < a } \phi_{b\,a} f_b \quad ( a=1,2,\dots,d ) \notag \end{align} Where the coefficients $ \phi_{b\,a} $ are recursively defined as follows: \begin{align} \phi_{b\,a} & = \frac{1}{ \xi^a(w) - \xi^b(w) } ( m_{b\,a}(w) + \sum_{b < c < a}m_{b\,c}(w)\phi_{c\,a}), \notag \\ \text{here} \quad \xi^a(w) & := \sum_{i=1}^{{\cal N}}w^{i-1}\xi_i^a \,, \qquad m_{a\,b}(w) \; := \; \sum_{i=1}^{{\cal N}}w^{i-1} m_i^{a\,b}\,. \notag \end{align} $ w \in {\Bbb C} $ and $ \phi_{b\,a} $ does not depend on w. \end{lemma} \mbox{} \noindent {\bf 6.} The joint characteristic number spectrum $ \{ \zeta_i^{\l}(\sigma) = p^{\l_{\sigma_i}}q^{l_{\sigma_i}} \}\quad ((i=1,\dots,N) ; \; $ \mbox{} $ \l \in \Lambda_N ; \; \sigma \in \S ) $ of the operators $ Y_i \quad (i=1,\dots,N) $ is explicitely multiplicity-free.The operators $ g_{i,j} \quad ( i,j = 1,2,\dots,N-1 ) $ preserve the finite-dimensional subspaces of $ {\Bbb C}[z_1,\dots,z_N] $ formed by homogeneous polynomials of any total degree. Therefore $ g_{i,j} \quad ( i,j = 1,2,\dots,N-1 ) $ and consequently $ Y_i \quad (i=1,\dots,N) $ are direct sums of finite dimensional operators. So we can apply the result of Lemma 1 and arrive at the following proposition: \begin{prop} There exist polynomials $ \Phi^{\l}_{\sigma}\quad (\l \in \Lambda_N ; \; \sigma \in \S ) $ s.t.: \begin{align} {\Bbb C}[z_1,\dots,z_N] & = \bigoplus_{\l \in \Lambda_N}E^{\l} \; , \; E^{\l}\::=\;\oplus_{\sigma \in \S} {\Bbb C}\Phi^{\l}_{\sigma} \, , \tag{i} \\ Y_i \Phi^{\l}_{\sigma} & = \zeta_i^{\l}(\sigma) \Phi^{\l}_{\sigma} \quad (i=1,\dots,N) , \tag{ii} \\ \Delta(u) \Phi^{\l}_{\sigma} & = \prod_{i=1}^N (1 + up^{\l_i}q^{l_i}) \Phi^{\l}_{\sigma} \,, \tag{ii'} \\ \Phi^{\l}_{\sigma} & = z^{\l_{\sigma}} + \text{{\em``s.m.''}}. \tag{iii} \end{align} \end{prop} \subsection{Action of the Hecke Algebra in the eigenspaces $E^{\l}\quad ( \l \in \Lambda_N )$ of the operator $\Delta (u)$.} \noindent {\bf 1.} Fix any $ \l \in \Lambda_N $ . Since $[\widehat{H_N(q)}\:,\: \Delta (u)]\:=\: 0$ with the action of $\widehat{H_N(q)}$ defined in {\bf 1.1.1}, we have:\begin{equation*} \widehat{H_N(q)}: \: E^{\l} \rightarrow E^{\l} \end{equation*} The affine generators of $ \widehat{H_N(q)} $ act in $ E^{\l} $ as given by (ii) in Proposition 2. In this section we find the action of the finite-dimensional Hecke Algebra generated by $ g_{i,i+1} \quad (i=1,...,N-1) $ on the polynomials $\Phi_{\sigma}^{\l} \quad ( \sigma \in \S ) $ forming a basis in $ E^{\l} \quad ( \l \in \Lambda_N ). $ \noindent {\bf 2.} Introduce operators $ T_{i,i+1}\; \in \; \widehat{H_N(q)} \quad ( i=1,\dots,N-1 ) $ : \begin{equation} T_{i,i+1} := g_{i,i+1} (Y_{i+1} - Y_i ) - (q-q^{-1}) Y_{i+1} \quad (i=1,\dots,N-1). \end{equation} These operators satisfy the following relations: \begin{align} T_{i,i+1}Y_i \; = \; Y_{i+1} T_{i,i+1}\:, & \quad T_{i,i+1}Y_{i+1} \; = \; Y_i T_{i,i+1} \quad (i=1,\dots,N-1) \\ \mbox{} [ T_{i,i+1}, Y_j ]& = 0 \qquad ( j \neq i,i+1) \end{align} Since $ T_{i,i+1}\; \in \; \widehat{H_N(q)} $, we have: $ T_{i,i+1}: E^{\l} \rightarrow E^{\l} \quad ( i=1,\dots,N-1 ). $ \mbox{} \noindent {\bf 3.} Consider the vector $ T_{i,i+1}\Phi_{\sigma}^{\l} \in E^{\l} \quad ( \sigma \in \S \; , \; i \in {1,2,\dots,N-1} ). $ Due to (1.1.43,44): \begin{eqnarray*} Y_k T_{i,i+1}\Phi_{\sigma}^{\l} & = & \xi_k^{\l}((i,i+1)\sigma) T_{i,i+1}\Phi_{\sigma}^{\l} \quad ( k = 1,\dots,N ). \end{eqnarray*} Since the spectrum of $ Y_k \quad ( k = 1,\dots,N ) $ on $ E^{\l}$ does not contain $ \xi_k^{\l}(\sigma) $ s.t. $ \sigma \not\in \S $, we have: \begin{align} T_{i,i+1}\Phi_{\sigma}^{\l} & = 0 \quad \text{when $ \sigma \in \S\,, (i,i+1)\sigma \not\in \S $} \quad (i=1,\dots,N-1). \end{align} Since the joint spectrum of $ Y_i \; (i=1,\dots,N) $ on $ E^{\l}$ is multiplicity-free, we have: \begin{align} T_{i,i+1}\Phi_{\sigma}^{\l} & = \tau_{i,i+1}^{\l}(\sigma) \Phi_{(i,i+1)\sigma}^{\l} \quad \text{when $ \sigma , (i,i+1)\sigma \in \S $} \quad (i=1,\dots,N-1). \end{align} Where $\tau_{i,i+1}^{\l}(\sigma)$ is a coefficient. Let $ \sigma \in \S $ and $ i $ be s.t. $ (i,i+1)\sigma \not\in \S $. Then $ \l_{\sigma_i} = \l_{\sigma_{i+1}} \;,\; \sigma_{i+1} = \sigma_i + 1 $, and (1.1.45) gives: \begin{align} (g_{i,i+1} - q )\Phi_{\sigma}^{\l} & = 0 . \end{align} Let $ \sigma \in \S $ and $ i $ be s.t. $ (i,i+1)\sigma \in \S $. Then we can recast (1.1.46) as follows: \begin{align} g_{i,i+1}\Phi_{\sigma}^{\l} & = \frac{(q-q^{-1})\zeta_{i+1}^{\l}(\sigma)}{\zeta_{i+1}^{\l}(\sigma) -\zeta_i^{\l}(\sigma)}\Phi_{\sigma}^{\l} + \frac{ \tau_{i,i+1}^{\l}(\sigma)}{\zeta_{i+1}^{\l}(\sigma)-\zeta_i^{\l}(\sigma)}\Phi_{(i,i+1)\sigma}^{\l}\, . \end{align} \noindent In order to find $\tau_{i,i+1}^{\l}(\sigma)$ we shall equate coefficients standing in front of monomials $ z^{\l_{\sigma}} \, ,\; z^{\l_{(i\,i+1)\sigma}} $ in the both sides of eq. (1.1.48). Recall that $ \sigma ,\, (i,i+1)\sigma \in \S $ entails in particular $ \l_{\sigma_i} \neq \l_{\sigma_{i+1}} $. \noindent Let $ \l_{\sigma_i} > \l_{\sigma_{i+1}}\,\Rightarrow\, z^{\l_{\sigma}} > z^{\l_{(i\,i+1)\sigma}} $. \noindent According to (1.1.38) the monomial $ z^{\l_{\sigma}} $ which is the maximal monomial in $\Phi_{\sigma}^{\l}$ appears in $g_{i,i+1}\Phi_{\sigma}^{\l}$ from two sources: from $\giz^{\l_{\sigma}}$ and from $\giz^{\l_{(i\,i+1)\sigma}}$. Denote by $x$ the coefficient at the monomial $z^{\l_{(i\,i+1)\sigma}}$ in $\Phi_{\sigma}^{\l}$. \noindent In the RHS of (1.1.48) $z^{\l_{\sigma}}$ appears as the maximal monomial in $\Phi_{\sigma}^{\l}$ and does not appear in $\Phi_{(i,i+1)\sigma}^{\l}$. Using (1.1.38) to compute the contributions from $g_{i,i+1}\Phi$ we equate the coefficients in front of $z^{\l_{\sigma}}$ in the both sides of (1.1.48): \begin{align} \d + q^{-1}x & = \frac{\d\zeta_{i+1}^{\l}(\sigma)}{\zeta_{i+1}^{\l}(\sigma) - \zeta_i^{\l}(\sigma)} . \end{align} Computing the contribution from $z^{\l_{(i\,i+1)\sigma}}$ we find that in the LHS of (1.1.48) this monomial appears only in $\giz^{\l_{\sigma}}$, while in the RHS it appears with coefficient $x$ in $\Phi_{\sigma}^{\l}$ and as the maximal monomial in $\Phi_{(i,i+1)\sigma}^{\l}$. Equating the coefficients we get: \begin{align} q & = \frac{\d\zeta_{i+1}^{\l}(\sigma)}{\zeta_{i+1}^{\l}(\sigma) - \zeta_i^{\l}(\sigma)} x + \frac{\tau_{i,i+1}^{\l}(\sigma)}{\zeta_{i+1}^{\l}(\sigma) - \zeta_i^{\l}(\sigma)} \,. \end{align} Combining (1.1.49) and (1.1.50) we obtain: \begin{align} \tau_{i,i+1}^{\l}(\sigma) & = q\frac{(q^{-1}\zeta_{i+1}^{\l}(\sigma) - q\zeta_i^{\l}(\sigma))(q\zeta_{i+1}^{\l}(\sigma) - q^{-1}\zeta_i^{\l}(\sigma))}{\zeta_{i+1}^{\l}(\sigma) - \zeta_i^{\l}(\sigma)} \quad ( \l_{\sigma_i} > \l_{\sigma_{i+1}}) \end{align} \mbox{} \noindent Let $ \l_{\sigma_i} < \l_{\sigma_{i+1}}\,\Rightarrow\, z^{\l_{\sigma}} < z^{\l_{(i\,i+1)\sigma}} $. Equate the coefficients in front of the monomial $z^{\l_{(i\,i+1)\sigma}}$ in the both sides of (1.1.48). \noindent In the LHS the contribution comes only from $\giz^{\l_{\sigma}}$. In the RHS only $\Phi_{(i,i+1)\sigma}^{\l}$ contributes $z^{\l_{(i\,i+1)\sigma}}$ as its maximal monomial. Application of (1.1.38) leads to: \begin{align} q^{-1} & = \frac{\tau_{i,i+1}^{\l}(\sigma)}{\zeta_{i+1}^{\l}(\sigma) - \zeta_i^{\l}(\sigma)} \quad ( \l_{\sigma_i} < \l_{\sigma_{i+1}}). \end{align} \mbox{} \noindent {\bf 4.} To summarize, we have obtained the following proposition: \begin{prop} Let $ \sigma \in \S \quad (\l \in \Lambda_N)$, then $H_N(q)$ acts in $E^{\l}\:=\:{\Bbb C}\{\Phi_{\sigma}^{\l} \}_{\sigma \in \S}$ as follows: \begin{equation} g_{i,i+1}\Phi_{\sigma}^{\l} = \frac{(q-q^{-1})\zeta_{i+1}^{\l}(\sigma)}{\zeta_{i+1}^{\l}(\sigma) -\zeta_i^{\l}(\sigma)}\Phi_{\sigma}^{\l} + \begin{cases} q\frac{(q^{-1}\zeta_{i+1}^{\l}(\sigma) - q\zeta_i^{\l}(\sigma))(q\zeta_{i+1}^{\l}(\sigma) - q^{-1}\zeta_i^{\l}(\sigma))}{(\zeta_{i+1}^{\l}(\sigma) - \zeta_i^{\l}(\sigma))(\zeta_{i+1}^{\l}(\sigma) - \zeta_i^{\l}(\sigma))}\Phi_{(i,i+1)\sigma}^{\l} & \\ \text{{\em when} $\l_{\sigma_i} > \l_{\sigma_{i+1}}\,\Rightarrow\, (i,i+1)\sigma \in \S $}, & \\ 0 \quad \text{{\em when} $\l_{\sigma_i} = \l_{\sigma_{i+1}}\,\Leftrightarrow\, (i,i+1)\sigma \not\in \S $}, & \\ q^{-1}\Phi_{(i,i+1)\sigma}^{\l} & \\ \text{{\em when} $\l_{\sigma_i} < \l_{\sigma_{i+1}}\,\Rightarrow\, (i,i+1)\sigma \in \S $}. & \end{cases} \end{equation} \end{prop} \section{The limit $p \rightarrow 1$ of the hierarchy of Dynamical Models.} \subsection{Few facts about Macdonald operators.} {\bf 1.}The Macdonald operators $ D_N^n(p,t) \quad (n=0,\dots,N)$ [M] act in the subspace of ${\Bbb C}[z_1,\dots,z_N]$ formed by symmetric polynomials. In notation of [JKKMP] these operators are defined as follows: \begin{equation} D_N^n(p,t) := t^{n(n-1)/2}\sum_{I_n}\prod\begin{Sb} i\in I_n \\ j\not\in I_n \end{Sb} \frac{tz_i - z_j}{z_i - z_j}\prod_{k\in I_n} p^{D_k} \quad (n=0,\dots,N), \end{equation} where the summation is over all subsets $ I_n $ of $\{1,2,\dots,N \}$ which contain $n$ elements. Using the formula: \begin{equation*} det\left\| \frac{(t-1)w_i}{tw_i-w_j}\right\|_{1\leq i,j\leq m} = \quad t^{m(m-1)/2}\prod_{1\leq i\neq j\leq m}\frac{w_i-w_j}{tw_i-w_j}, \end{equation*} where $w_i \quad (i=1,\dots,m)$ are numbers; we can rewrite the definition of the operators $ D_N^n(p,t) \quad (n=0,\dots,N)$ in another form: \begin{equation} D_N^n(p,t) = \sum_{I_n}detA_{I_n}(t)\prod_{k\in I_n} p^{D_k} \quad (n=0,\dots,N), \end{equation} where $A_{I_n}(t)$ is a submatrix of the matrix: \begin{equation} A(t) = \| A_{ij}(t)\|_{1\leq i,j\leq N}\;,\;A_{ij}(t):=\frac{(t-1)z_i}{tz_i - z_j} \prod\begin{Sb} 1\leq k\leq N \\ k\neq i \end{Sb}\frac{tz_i - z_k}{z_i - z_k} ; \end{equation} which is defined as follows:$ A_{I_n}(t):=\| A_{ij}(t)\|_{i,j\in I_n}.$ The Macdonald polynomials $P_{\l}(p,t) \quad (\l \in \Lambda_N)$ are eigenfunctions of the operators $ D_N^n(p,t) \quad (n=0,\dots,N)$: \begin{equation} D(v;p,t)P_{\l}(p,t) = \prod_{i=1}^N (1 + t^{N-i}p^{\l_i}v)\, P_{\l}(p,t) \quad (\l \in \Lambda_N), \end{equation} here $D(v;p,t)$ is the generating function of Macdonald operators: \begin{equation*} D(v;p,t):= \sum_{n=0}^N v^n D_N^n(p,t) \;. \end{equation*} \noindent {\bf 2.} In the limit $p \rightarrow 1$ one finds [M]: \begin{equation} P_{\l}(p,t) = e_{\l '} + O(p-1) \quad (\l \in \Lambda_N), \end{equation} where $\l '$ is the conjugate partition of $\l$ and for a partition $\pi \: : (\pi_1 \geq \pi_2 \geq \dots ) \; , \; (\pi_1 \leq N) \quad e_{\pi}:= e_{\pi_1}e_{\pi_2}\ldots \quad ;$ where $ e_r $ is the elementary symmetric polynomial: \begin{equation*} e_r := \sum_{1\leq i_1 \leq\dots\leq i_r\leq N} z_{i_1}z_{i_2}\ldots z_{i_r}. \end{equation*} Consider the limit $p \rightarrow 1$ of the generating function $D(v;p,t)$: \begin{equation} D(v;p,t) \overset{p \rightarrow 1}{=} D_0(v;t) + (p-1) D_1(v;t) + O((p-1)^2). \end{equation} {}From (2.2.1,.4,.5) it follows that $D_0(v;t)$ is a multiplication by a constant: \begin{equation} D_0(v;t) = \prod_{i=1}^N (1 + t^{N-i}v) \;. \end{equation} The first-order term $D_1(v;t)$ is a differential operator: \begin{equation} D_1(v;t) = \sum_{i=1}^N \left(\sum_{n=1}^N v^n \sum_{I_n: i\in I_n} detA_{I_n}(t)\right)D_i \end{equation} Expanding the eq. (2.2.4) up to the first order in $ p-1 $ and using (2.2.5,.7 ) we get: \begin{equation} D_1(v;t)e_{\l '} = \left(v\sum_{j=1}^N\prod\begin{Sb} 1\leq k\leq N \\ k\neq j \end{Sb} (1 + v t^{N-k}) t^{N-j}\l_j\right)\,e_{\l '} \quad ( \l \in \Lambda_N ). \end{equation} \subsection{Taking the limit $p \rightarrow 1$ in the hierarchy of Dynamical Models} {\bf 1.} The following fact was established in the paper [JKKMP]. Let $S$ be any symmetric polynomial $ ( S \in {\Bbb C}[z_1,\dots,z_N] )$. The action of the operator $\Delta (u)$ (cf. {\bf 1.1}) on such $S$ coincides with the action of the generating function of Macdonald operators: \begin{equation} \Delta (u) S = D(q^{N-1}u;p,q^{-2})S \;. \end{equation} There is another connection between $\Delta (u)$ and Macdonald operators. For an operator $O$ which is a function of operators $ D_i , z_i \; (i=1,\dots,N) , K_{i,i+1} $ $ (i = 1,\dots,N-1) $ introduce a normal ordering $:\; :$. The normal ordering is described as follows: in $O$ bring all the operators $ D_i $ to the right {\em without} taking commutators between $ D_i $ and $ z_j $, but taking into account the commutation relations between $ D_i $ and $ K_{j,j+1} $. For instance: \begin{equation*} :\; p^{D_1}\, \frac{q^{-1}z_1 - qz_2}{z_1-z_2}(K_{12} - 1)\; : \;= \frac{q^{-1}z_1 - qz_2}{z_1-z_2}(K_{12} \,p^{D_2}- p^{D_1}) . \end{equation*} \noindent We formulate the following Lemma: \begin{lemma} \begin{equation} :\; \Delta (u) \; : \; = D(q^{N-1}u;p,q^{-2}) \; . \end{equation} \end{lemma} \begin{pf} To facilitate the proof we introduce an extension of the algebra generated by $ z_1,z_2,\dots,z_N $ and $ K_{1,2},K_{2,3},\dots,K_{N-1,N}$ by symbols $ \xi_1,\xi_2,\dots,\xi_N $. These symbols are defined by the commutation relations \begin{align*} [\xi_i,\xi_j] & = 0 \; , \; [\xi_i,z_j] = 0 \qquad ( i,j \in \{1,2,\dots,N \}), \\ K_{i,j}\xi_j & = \xi_i K_{i,j}\;,\; [K_{i,j},\xi_k] = 0 \quad (k\neq i,j) \qquad ( i\neq j \in \{1,2,\dots,N \}). \end{align*} One can take $\xi_i := f(z_i) $, where $f$ is any function of one variable, as a realization for $ \xi_i$. \\ Together with $ g_{i,i+1} $ the operators \begin{equation*} Y_i(\xi) := g_{i,i+1}^{-1}K_{i,i+1}\dots g_{i,N}^{-1}K_{i,N}\xi_i K_{1,i}g_{1,i}\dots K_{i-1,i}g_{i-1,i} \end{equation*} still satisfy the defining relations of $ \widehat{H_N(q)}$. \\ This implies in particular that for the operator \begin{equation*} \Delta(u;\xi) := \prod_{i=1}^N(1 + u Y_i(\xi)) \end{equation*} we have \begin{equation} [\Delta(u;\xi),g_{i,i+1}] = 0 \qquad (i=1,\dots,N-1). \tag{i} \end{equation} Using the commutation relations with $ z_1,z_2,\dots,z_N $ and $ K_{i,j} $ we can bring the symbols $ \xi_i $ to the right of all expressions in $\Delta(u;\xi)$. Denote $\Delta(u;\xi)$ with all $ \xi_i $ brought to the right by $\Delta(u;\xi)'.$ \\ We have \begin{equation*} : \Delta(u) : = \Delta(u;\xi)' |_{\xi_i \rightarrow p^{D_i}}. \end{equation*} Therefore in order to prove the statement of the lemma we compute the coefficients standing in front of monomials $ \xi_{i_1}\xi_{i_2}\dots \xi_{i_n} \quad (1\leq i_1 \leq i_2 \leq \dots \leq i_n \leq N)$ in the symmetric functions \begin{equation*} \Delta(\xi)^{(n) } := \sum_{N \geq k_1 > k_2 > \dots > k_n \geq 1 } Y_{k_1}(\xi)Y_{k_2}(\xi)\dots Y_{k_n}(\xi) \quad ( 1 \leq n \leq N ). \end{equation*} With notation of {\em {\bf 1.1.1}} we have \begin{align*} r_{i,j} & := K_{i,j}g_{i,j} = a_{j,i} + b_{j,i}K_{i,j} , \\ r_{i,j}^{-1} & = a_{i,j} - b_{j,i}K_{i,j}, \\ Y_i(\xi) & = r_{i,i+1}^{-1}\dots r_{i,N}^{-1}\xi_i r_{1,i}\dots r_{i-1,i}. \end{align*} \mbox{} \noindent Let us compute the terms in $\Delta(\xi)^{(n)'} \quad ( 1\leq n \leq N)$ which contain symbols $ \xi_1 ,\xi_2, \dots ,\xi_n $ only. By inspection we find that such terms can appear only in $ Y_n(\xi)Y_{n-1}(\xi)\dots Y_1(\xi) $. Furthermore the relevant contributions from the individual factors in the last expression are \begin{align*} Y_1(\xi) & \rightarrow r_{1,2}^{-1}\dots r_{1,n}^{-1}a_{1,n+1}\dots a_{1,N} \xi_1 , \\ Y_2(\xi) & \rightarrow r_{2,3}^{-1}\dots r_{2,n}^{-1}a_{2,n+1}\dots a_{2,N} \xi_2 r_{1,2}, \\ & \vdots \\ Y_k(\xi) & \rightarrow r_{k,k+1}^{-1}\dots r_{k,n}^{-1}a_{k,n+1}\dots a_{k,N} \xi_k r_{1,k}\dots r_{k-1,k}, \\ & \vdots \\ Y_{n-1}(\xi) & \rightarrow r_{n-1,n}^{-1}a_{n-1,n+1}\dots a_{n-1,N} \xi_{n-1} r_{1,n-1}\dots r_{n-2,n-1}, \\ Y_n(\xi) & \rightarrow a_{n,n+1}\dots a_{n,N} \xi_n r_{1,n}\dots r_{n-1,n}. \end{align*} Multiplying these contributions we find that there is only one term in $\Delta(\xi)^{(n)'}$ which contains $ \xi_1 ,\xi_2, \dots ,\xi_n $ only; and this term is \begin{multline*} (a_{1,n+1}\dots a_{1,N})(a_{2,n+1}\dots a_{2,N})\ldots (a_{n,n+1}\dots a_{n,N})\xi_1\xi_2\dots\xi_n . \end{multline*} Next we use the Hecke-invariance relation {\em (i) } and find \begin{equation*} \Delta(\xi)^{(n)} = \sum_{I_n} \left(\prod\begin{Sb} i \in I_n \\ j \not\in I_n \end{Sb} a_{i,j}\right) \prod_{i \in I_n } \xi_i. \end{equation*} Where the summation is over all $n$-element subsets of $\{1,2,\dots,N \}$. Comparing this expression with {\em (2.2.1)} we obtain the statement of the lemma. \end{pf} \mbox{} \noindent {\bf 2.} Let us take the limit $p \rightarrow 1$ in the operators $ \Delta (u) , T_a(u), Y_i \quad (i=1,\dots,N).$ Expanding around $ p = 1 $ and keeping the first two terms of the expansion we write: \begin{align} Y_i &\overset{p \rightarrow 1}{=} y_i + (p-1)x_i + O((p-1)^2)\quad ((i=1,\dots,N)), \\ \Delta (u) &\overset{p \rightarrow 1}{=} \Delta_0(u) + (p-1)\Delta_1(u)+ O((p-1)^2)\:, \\ T_a(u) &\overset{p \rightarrow 1}{=} T_a^0(u) + (p-1) T_a^1(u)+ O((p-1)^2)\: . \end{align} Here we introduced the operators: \begin{align} y_i & := g_{i,i+1}^{-1}K_{i,i+1}\dots g_{i,N}^{-1}K_{i,N} K_{1,i}g_{1,i}\dots K_{i-1,i}g_{i-1,i} \quad ((i=1,\dots,N)) , \\ x_i & := g_{i,i+1}^{-1}K_{i,i+1}\dots g_{i,N}^{-1}K_{i,N}D_i K_{1,i}g_{1,i}\dots K_{i-1,i}g_{i-1,i} \quad ((i=1,\dots,N)) , \\ & \Delta_0(u) := \prod_{i=1}^N(1 + uy_i) \;, \\ & \Delta_1(u) := u\sum_{j=1}^N \prod_{1\leq i<j}(1 + uy_i)\, x_j\prod_{j<k\leq N}(1 + uy_k)\;,\\ T_a^0(u) & := L_{a1}(uy_1)L_{a2}(uy_2)\ldots L_{aN}(uy_N) \quad \in End{\Bbb C}[z_1,\dots,z_N]\otimes H\;. \end{align} The operators $y_i , g_{i,i+1}$ satisfy the relations (cf. {\bf 1.1}) of Affine Hecke Algebra and $ T_a^0(u)$ defines a representation of $U$. \noindent {\bf 3.} From Lemma 2 it follows that the operators $ \Delta_0(u)\, , \, \Delta_1(u) $ have a rather special form. We have the proposition: \begin{prop} The following statements hold: \noindent Let $D_0(v;t)\, , \; D_1(v;t)$ be those defined in {\em (2.2.6,.8)}; then: \begin{align} \Delta_0(u) & = D_0(q^{N-1}u;q^{-2}) \, = \, \prod_{i=1}^N (1 + u q^{2i-N-1})I \, . \tag{i} \end{align} I.e. $\Delta_0(u)$ is a multiplication by a constant. \begin{align} \Delta_1(u) & = D_1(q^{N-1}u;q^{-2}) + \Xi(u) \,. \tag{ii} \end{align} Where operator $ \Xi(u)$ is a function of operators $ z_i \,,\, K_{i,j} \quad (i,j = 1,\dots, N) $ only ( and {\em not} of $ D_i $). \end{prop} \mbox{} \noindent Let $ \Delta (u)_1 := \sum_{i=1}^{N}u^i \Delta_1^{(i)} $ and $ \Xi(u) := \sum_{i=1}^{N}u^i \Xi^{(i)} $. We have computed explicit expressions for the operators $ \Delta_1^{(N)} $ and $ \Delta_1^{(1)} $. In notation of section {\bf 1.1} one has: \begin{align} \Delta_1^{(N)} & = D_1 + D_2 + \dots + D_N \, , \\ \Delta_1^{(1)} & = \sum_{i=1}^N \left( \prod\begin{Sb} 1\leq k\leq N \\ k\neq i \end{Sb} a_{i,k} \right)D_i \, + \, \Xi^{(1)}\, \end{align} where: \begin{multline} \Xi^{(1)} = \\ \sum_{M=2}^N \frac{(-1)^M}{\d}\sum_{N\geq i_M > \dots >i_1\geq 1} {\cal A}_{i_M,i_{M-1},\dots,i_1}{\cal B}_{i_M,i_{M-1},\dots,i_1} K_{i_M,i_{M-1}}\dots K_{i_2,i_1} \; + \; \varphi^{(1)} \:, \notag \\ {\cal A}_{i_M,i_{M-1},\dots,i_1} = \left(\prod_{i_1<f<i_2} a_{i_1,f}\right)\left(\prod_{i_2<f<i_3} a_{i_2,f}\right)\ldots \left(\prod_{i_M<f<N+i_1} a_{i_M,f\pmod{N}}\right) \:, \\ {\cal B}_{i_M,i_{M-1},\dots,i_1} = b_{i_M,i_{M-1}}b_{i_M-1,i_{M-2}}\ldots b_{i_2,i_{1}}b_{i_1,i_M} \:, \\ \varphi^{(1)} = -\sum_{1\leq k< i\leq N}\frac{a_{i,i+1}\dots a_{i,N} a_{i,1}\dots a_{i,k-1}a_{i,k+1}\dots a_{i,i-1} b_{k,i} b_{i,k}}{\d} \: . \notag \end{multline} \mbox{} \noindent {\bf 4.} Let us fix the notation: \begin{equation} {\cal D}(u) := D_1(q^{N-1}u;q^{-2}) = \sum_{i=1}^N \theta_i(u)\,D_i\, . \end{equation} Where according to (2.2.8) : \begin{equation} \theta_i(u) := \sum_{n=1}^N u^n q^{n(N-1)}\sum_{I_n : i\in I_n} det A_{I_n}(q^{-2}) \quad ((i=1,\dots,N)). \end{equation} (Cf. sec. {\bf 2.1} for the definition of $A_{I_n}(t)$). The functions $\theta_i(u)\quad ((i=1,\dots,N))$ can be written in the following form: \begin{equation} \theta_i(u) = \frac{\partial}{\partial \gamma_i}det(I + u\Gamma q^{N-1} A(q^{-2}))|_{\Gamma = I} \, , \end{equation} where we have introduced an auxiliary matrix: $ \Gamma := {\mathrm {diag}}\{\gamma_1,\ldots, \gamma_N \} .$ Using this representation we compute $\theta_i(u)\quad ((i=1,\dots,N))$ at the point $ z_1 = \omega^1,\dots,z_N = \omega^N $, where $\omega := {\mathrm {exp}}(2\pi i/N).$ The computation yields: \begin{gather} \theta_i(u)|_{z_1=\omega^1,\dots,z_N=\omega^N} = \frac{1}{N}\prod_{k=1}^N (1+uq^{l_k})\,\sum_{n=1}^N \frac{uq^{l_n}}{1+uq^{l_n}} \; \equiv \theta(u) \quad (i=1,\dots,N). \\ l_i := 2i - N - 1 \notag \end{gather} Thus the point $ z_1 = \omega^1,\dots,z_N = \omega^N $ (or any point obtained from it by a permutation of coordinates) is special in that at this point all the $\{z_i\}$-dependent coefficients $\theta_i(u)$ of the first-order differential operator ${\cal D}(u)$ become equal one to another. \section{Definition of the Hierarchy of Integrable, $U$-invariant Spin Models} \subsection{Preliminaries} {\bf 1.} Let us expand the relations (1.1.34,.35) around the point $p=1$ using the definitions (2.2.12-.14) and the fact that $\Delta_0(u)$ is a constant. For $\Delta_1(u),\,T_a^0(v) \in End({\Bbb C}[z_1,\dots,z_N]\otimes H)$ we obtain: \begin{align} [\Delta (u)_1,T_a^0(v)] & = 0 \; , \\ \mbox{} [\Delta_1(u),\Delta_1(v)] & = 0. \end{align} Expanding the relations: \begin{equation} [\Delta (u),Y_i] = 0 \quad (i=1,\dots,N),\quad [\Delta (u),g_{i,i+1}] = 0 \quad (i=1,\dots,N-1), \end{equation} we get: \begin{equation} [\Delta_1(u),y_i] = 0 \quad (i=1,\dots,N),\quad [\Delta_1(u),g_{i,i+1}] = 0 \quad (i=1,\dots,N-1). \end{equation} Due to (3.3.4) and the fact that $y_i \quad (i=1,\dots,N) ,\; g_{i,i+1} \quad (i=1,\dots,N-1)$ satisfy the Affine Hecke Algebra relations, the operators $\Delta_1(u), T_a^0(v)$ act in the ``bosonic'' subspace ${\cal B}$ (cf. 1.1.36 and 3.3.7). \mbox{} \noindent {\bf 2.} Introduce several definitions. Let \begin{align} {\cal R} & := {\Bbb C} [\{\frac{1}{z_i - z_j}\}_{1\leq i\neq j\leq N}\, ,\{\frac{1}{z_i - q^{2}z_j}\}_{1\leq i\neq j\leq N}\,,\: z_1,\dots,z_N ]\otimes H. \end{align} Let ${\cal P}$ and ${\cal B}$ be the subspaces of $ {\cal R}$: \begin{align} {\cal P} & := {\Bbb C}[z_1,\dots,z_N]\otimes H \, , \\ {\cal B} & := \{b\in {\cal P} | (g_{i,i+1} - t_{i,i+1})b = 0 \quad (i=1,\dots,N-1)\}. \end{align} For $v_1,\dots,v_N \, \in {\Bbb C} $ such that $ v_i \neq v_j\,,\,q^{2} v_j \quad (i\neq j\,; i,j =1,\dots,N)$ define the evaluation map: $ Ev(v): {\cal R} \mapsto H $ by taking values of rational functions at the point $ z_1 = v_1,\dots, z_N = v_N.\; ( \Leftrightarrow z = v ) $. \noindent For any $ O \in End({\cal R}) $ define an operator $ \hat{O} \in End({\cal R}) $ by the rule [BGHP]: Using the commutation relations bring all the permutation operators $K_{i,j}\quad (i,j \in \{1,2,\dots,N \})$ to the right of an expression in $O$; replace the rightmost of $K_{i,j}$ using the substitution: \begin{equation} g_{i,i+1} \rightarrow t_{i,i+1} \Rightarrow K_{i,i+1} \rightarrow \frac{z_it_{i,i+1}^{-1} - z_{i+1}t_{i,i+1}}{q^{-1}z_i - qz_{i+1}} \quad (i=1,\dots,N-1). \end{equation} Repeat the procedure until there are no operators $K_{i,j}$ left. The result is $ \hat{O}$. In what follows we adopt the following notational convention: if ${\cal L}$ is a linear space and $A,B$ are linear operators defined on ${\cal L}$, we write: \begin{gather*} A{\cal L} = B{\cal L}\quad \text{meaning } \quad Al = Bl \quad \forall l \in {\cal L}. \end{gather*} In particular for $O , \hat{O} \quad \in End({\cal R})$ defined above we have: \begin{equation} O{\cal B} = \hat{O}{\cal B}. \end{equation} Let $O,O' \in End({\cal P})$ be s.t.: \begin{gather} [O,O']{\cal P} = 0 \quad \text{and}\quad O ,O': {\cal B}\mapsto {\cal B}, \end{gather} then \begin{equation} [\hat{O},\hat{O'}]{\cal B} = 0. \end{equation} \noindent {\bf 3.}With notation of (2.2.20,.23,.26) let us consider the following differential operator $\widetilde{{\cal D}(u)} \in End({\cal R})$: \begin{equation} \widetilde{{\cal D}(u)} := {\cal D}(u) - \theta(u)\Delta_1^{(N)}. \end{equation} Let $Ev(\omega)$ be the evaluation map $Ev(v)$ taken at the special point $ v_1 = \omega^1,\dots,v_N = \omega^N \quad ( \Leftrightarrow v = \omega ) $. Then in virtue of (2.2.26) we obtain the following property of $ \widetilde{{\cal D}(u)} $: \begin{equation} Ev(\omega)\widetilde{{\cal D}(u)}{\cal R} = 0. \end{equation} Let us introduce the modified generating function $ \widetilde{\Delta_1(u)} $ by subtracting the product of the constant $\theta(u)$ and the operator $\Delta_1^{(N)}$: \begin{equation} \widetilde{\Delta_1(u)} := \Delta_1(u) - \theta(u)\Delta_1^{(N)} = \widetilde{{\cal D}(u)} + \Xi(u) . \end{equation} Since $\Delta_1^{(N)}$ is a member of the hierarchy of commuting operators defined by $\Delta_1(u)$, the equations (3.3.1,.2),(3.3.4) still hold if we replace in these equations $\Delta_1(u)$ by $\widetilde{\Delta_1(u)}$: \begin{align} [\widetilde{\Delta_1(u)},\widetilde{\Delta_1(v)}]{\cal P} & = 0 , \\ [\widetilde{\Delta_1(u)}, T_a^0(v) ]{\cal P} & = 0 , \\ \widetilde{\Delta_1(u)} : {\cal B} & \mapsto {\cal B} . \end{align} \subsection{Definition of the hierarchy of Spin Models} {\bf 1.} Let $ H_{{\cal B}}(\omega)$ be the image of ${\cal B}$ under the action of the evaluation map $Ev(\omega)$: \begin{equation} Ev(\omega){\cal B} = H_{{\cal B}}(\omega) \subset H . \end{equation} Since $ \widehat{T_a^0(u)}\;,\,\widehat{\Xi(u)} $ do not depend on the diffrential operators $ D_i \; (i=1,\dots,N) $ and the operators of coordinate permutation, we can define the operators $T_a^0(u;\omega) , \Xi(u;\omega)$ as follows: \begin{equation} Ev(\omega)\widehat{T_a^0(u)} = T_a^0(u;\omega) Ev(\omega) , \quad Ev(\omega)\widehat{\Xi(u)} = \Xi(u;\omega) Ev(\omega). \end{equation} Applying $Ev(\omega)$ to the relations : \begin{align} \widehat{T_a^0(u)}:{\cal B} & \mapsto {\cal B} , \\ \widehat{\widetilde{\Delta_1(u)}}:{\cal B} & \mapsto {\cal B}, \end{align} and using (3.3.13) we get: \begin{align} T_a^0(u;\omega) : H_{{\cal B}}(\omega) & \mapsto H_{{\cal B}}(\omega) , \\ \Xi(u;\omega) : H_{{\cal B}}(\omega) & \mapsto H_{{\cal B}}(\omega). \end{align} \noindent {\bf 2.} Apply the evaluation map $Ev(\omega)$ to the relations (3.3.15,.16) taking (3.3.13) and (3.3.23) into account. As the result we find that the operator $\Xi(u;\omega)$ is a generating function of the commuting, $U$-invariant integrals of motion which are operators in $ H_{{\cal B}}(\omega)$: \begin{gather} [\Xi(u;\omega), \Xi(v;\omega) ] H_{{\cal B}}(\omega) = 0 \, , \\ [\Xi(u;\omega) , T_a^0(v;\omega) ] H_{{\cal B}}(\omega) = 0 , \\ \left(\bar{R}_{ab}(u/v)T_a^0(u;\omega) T_b^0(v;\omega) - T_b^0(v;\omega)T_a^0(u;\omega)\bar{R}_{ab}(u/v)\right) H_{{\cal B}}(\omega) = 0 . \end{gather} In sec. 6 we shall show that $H_{{\cal B}}(\omega)=H$. This completes the definition of the hierarchy $ \Xi(\omega)^{(1)},\dots,\Xi(\omega)^{(N-1)} \; ( \Xi(u;\omega) = \sum_{n=1}^{N-1} u^n \Xi^{(n)}(\omega) ) $ of Spin Models. \section{Eigenvalue spectrum of the operators $\Delta_1(u)\;, \protect\newline y_i \; (i=1,\dots,N) $} \subsection{Characteristic numbers and eigenvalues of $\Delta_1(u)\;,\;y_i $} {\bf 1.} To find the action of the operators $\Delta_1(u)\;,\;y_i \; (i=1,\dots,N)$ in the monomial basis of ${\Bbb C}[z_1,\dots,z_N]$ we can take the limit $p\rightarrow 1$ in the formulas of Proposition 1. This gives the following proposition: \begin{prop} The operators $\Delta_1(u)\;,\;y_i \; (i=1,\dots,N)$ are triangular in the monomial basis of ${\Bbb C}[z_1,\dots,z_N]$. The action of these operators on monomials is given by: \begin{align} y_iz^{\l_{\sigma}} & = q^{l_{\sigma_i}}z^{\l_{\sigma}} + \text{{\em``s.m''}}\quad (\l \in \Lambda_n, \sigma \in \S, \; (i=1,\dots,N) ), \tag{i} \\ \Duz^{\l_{\sigma}} & = \delta^{\l}(u)z^{\l_{\sigma}} + \text{{\em``s.m''}}\quad (\l \in \Lambda_N, \sigma \in \S ) \tag{ii}. \end{align} where $l_i := 2i - N -1 \quad (i=1,\dots,N) $ and \begin{equation*} \delta^{\l}(u) := u\sum_{j=1}^N\left(\prod\begin{Sb} 1\leq k\leq N \\ k\neq j \end{Sb}(1 + uq^{l_k})\right)q^{l_j}\l_j . \end{equation*} \end{prop} Since $\Delta_1(u)\;$ and $\;y_i \; (i=1,\dots,N)$ commute among themselves (cf. 3.3.4) and the joint spectrum of characteristic numbers of $\Delta_1(u)\;$ and $\;y_i \; (i=1,\dots,N)$ given by Proposition 5 is explicitely multiplicity-free, we apply Lemma 1 and claim that $\Delta_1(u)\;,\;y_i \; (i=1,\dots,N)$ are simultaneously diagonalizable: \begin{prop} There exist polynomials $\varphi_{\sigma}^{\l}\quad (\l \in \Lambda_N, \sigma \in \S )$ s.t.: \begin{align} {\Bbb C}[z_1,\dots,z_N] & = {\bigoplus}_{\l \in \Lambda_N}{\cal E}^{\l}\;,\;{\cal E}^{\l}:={\oplus}_{\sigma\in\S}{\Bbb C}\varphi_{\sigma}^{\l}\,,\tag{i}\\ y_i\varphi_{\sigma}^{\l} & = q^{l_{\sigma_i}}\varphi_{\sigma}^{\l} \quad (i=1,\dots,N) ,\tag{ii}\\ \Delta_1(u)\varphi_{\sigma}^{\l} & = \delta^{\l}(u)\varphi_{\sigma}^{\l}\,,\tag{ii'}\\ \varphi_{\sigma}^{\l} & = z^{\l_{\sigma}} + \text{{\em``s.m''}}\,. \end{align} \end{prop} \mbox{} \noindent {\bf 2.} Let us show that $ \varphi_{\sigma}^{\l} = {\lim}_{p\rightarrow 1} \Phi_{\sigma}^{\l}\quad (\l \in \Lambda_N, \sigma \in \S )$ where $\Phi_{\sigma}^{\l}\quad (\l \in \Lambda_N, \sigma \in \S )$ are eigenfunctions of the operators $Y_i\quad(i=1,\dots,N)$ (Cf. Proposition 2). Expanding $\Phi_{\sigma}^{\l}\quad (\l \in \Lambda_N, \sigma \in \S )$ around $p = 1$ let us write: \begin{equation} \Phi_{\sigma}^{\l} \overset{p\rightarrow 1}{=} (p - 1)^{s} \psi^{\l}_{\sigma} + O((p-1)^{s+1}), \end{equation} where $ \psi^{\l}_{\sigma}$ is a non-zero polynomial. Since $ \Phi_{\sigma}^{\l} = z^{\l_{\sigma}} + \text{{\em `` s.m.''}} $, we have: $s \leq 0$. Let us show that $s=0$. Suppose $s<0$. The satement (c) of Lemma 1 when applied to the eigenvectors $\Phi_{\sigma}^{\l}$ enables us to detect which coefficients in the decomposition of $\Phi_{\sigma}^{\l}$ into monomials are potentially singular in the limit $p \rightarrow 1$. The singularities may arise because of the presence of denominators of the form \begin{equation} \frac{1}{p^{\l_{\sigma_i}}q^{l_{\sigma_i}} - p^{\mu_{\sigma_i}}q^{l_{\sigma_i}}} \end{equation} where $\mu$ is a partition {\em smaller} than $\l$ and $\sigma \in \S , S^{\mu}_N $.Therefore if $s$ in (4.4.2) is negative, the maximal monomial in $ \psi^{\l}_{\sigma}$ is {\em smaller } than any of the monomials $z^{\l_{\sigma}} \quad (\sigma \in \S)$. Expanding the equations (ii),(ii') of the Proposition 2 around the point $p=1$ and taking into account that $\Delta_0(u) = \Delta (u) |_{p=1}$ is a constant, we arrive at the following equations: \begin{align*} \Delta_1(u)\psi^{\l}_{\sigma} & = \delta^{\l}(u)\psi^{\l}_{\sigma}\,, \\ y_i\psi^{\l}_{\sigma} & = q^{l_{\sigma_i}} \psi^{\l}_{\sigma}. \end{align*} Since the joint spectrum of the operators $\Delta_1(u)$ and $y_i \; (i=1,\dots,N)$ is multiplicity-free, we must have: \begin{equation*} \psi^{\l}_{\sigma} \propto \varphi_{\sigma}^{\l}\,. \end{equation*} This contradicts the observation that the maximal monomial of $\psi^{\l}_{\sigma}$ is smaller than any of the monomials $z^{\l_{\sigma}} \quad (\sigma \in \S)$. Thus $s=0$ therefore $ \psi^{\l}_{\sigma} = \varphi_{\sigma}^{\l} $ and consequently $ \varphi_{\sigma}^{\l} = {\lim}_{p\rightarrow 1}\Phi_{\sigma}^{\l}\quad (\l \in \Lambda_N, \sigma \in \S )$. \subsection{Action of the Hecke Algebra in the eigenspaces ${\cal E}^{\l}\;(\l \in \Lambda_N)$ of the operator $\Delta_1(u)$ .} {\bf 1.} According to (3.3.4) the Hecke Algebra $H_N(q)$ generated by $ g_{i,i+1} \; (i=1,\dots,N-1) $ acts in each eigenspace ${\cal E}^{\l}\;(\l \in \Lambda_N)$ of the operator $\Delta_1(u)$. To compute this action explicitely in the basis $\{ \varphi_{\sigma}^{\l} \}_{\sigma \in \S}$ we can either repeat almost word-by-word the derivation described in {\bf 1.5 } or take the limit $p \rightarrow 1$ in the result of Proposition 3. Either way we arrive at the following proposition: \begin{prop} Let $\sigma \in \S \quad (\l\in \Lambda_N)$, then $H_N(q)$ generated by $ g_{i,i+1} \; (i=1,\dots,N-1) $ acts in ${\cal E}^{\l} = {\Bbb C} \{ \varphi_{\sigma}^{\l} \}_{\sigma \in \S}$ as follows: \begin{equation} g_{i,i+1}\varphi_{\sigma}^{\l} = \frac{\d q^{l_{\sigma_{i+1}}}}{q^{l_{\sigma_{i+1}}}-q^{l_{\sigma_i}}}\varphi_{\sigma}^{\l} + \begin{cases} & q\frac{(q^{-1}q^{l_{\sigma_{i+1}}}-qq^{l_{\sigma_i}})(qq^{l_{\sigma_{i+1}}}-q^{-1}q^{l_{\sigma_i}})}{(q^{l_{\sigma_{i+1}}}-q^{l_{\sigma_i}})(q^{l_{\sigma_{i+1}}}-q^{l_{\sigma_i}})}\varphi_{(i,i+1)\sigma}^{\l} \\ & \text{{\em when} $ \l_{\sigma_i} > \l_{\sigma_{i+1}} \Rightarrow (i,i+1)\sigma \in \S $} , \\ & 0 \quad \text{{\em when} $\l_{\sigma_i} = \l_{\sigma_{i+1}} \Leftrightarrow (i,i+1)\sigma \not\in \S $}, \\ & q^{-1}\varphi_{(i,i+1)\sigma}^{\l} \\ & \text{{\em when} $ \l_{\sigma_i} < \l_{\sigma_{i+1}} \Rightarrow (i,i+1)\sigma \in \S $}. \end{cases} \end{equation} \end{prop} \subsection{``Motifs'' and associated partitions} {\bf 1.} Following [HHTBP,BPS] introduce the definition: \begin{df} Call a sequence of $M$ integers $(m_1,m_2,\dots,m_M)$ a {\em motif} iff: \begin{align} 1 \leq m_1 < m_2 < \ldots < m_M \leq N-1 \, , \tag{i} \\ m_{i+1} \geq m_i + 2 \quad (i=1,\dots,M-1). \tag{ii} \end{align} \end{df} With any motif $(m_1,m_2,\dots,m_M)$ associate a partition $\mu$: \begin{equation*} \mu = (M,\dots,\underset{m_1}{M},M-1,\dots,\underset{m_2}{M-1}, \; \ldots \; ,\underset{m_{M-1}+1}{1},\dots,\underset{m_M}{1},0,\dots,\underset{N}{0}) . \end{equation*} The subscripts in the last equation indicate positions of numbers in the partition. In what follows we shall identify motifs with the partitions they define. We shall indiscriminately use the notation $(m_1,m_2,\dots,m_M)$ for both a motif and the corresponding partition. Let $ {\frak M}_N $ be the set of all motifs for a fixed $N$. We use the same notation for the corresponding subset of all partitions. \mbox{} \noindent {\bf 2.} Let $(m_1,m_2,\dots,m_M)$ be a partition from the set $ {\frak M}_N $. We subdivide the set $S_N^{\m}$ (cf. {\bf 1.4 }) into disjoint subsets: \begin{df} For any subset $\{i_1,i_2,\dots ,i_L\} \subset \{1,2,\dots,M \} \quad (0\leq L\leq M) $ define $ S_{N,\mi}^{\m} \subset S_N^{\m} \subset S_N $ as follows: \begin{eqnarray*} \lefteqn{S_{N,(\emptyset)}^{(m_1,m_2,\dots,m_M)} := \{ {\mathrm {id}} \} , } \\ \lefteqn{\text{{\em for } $ 1 \leq L \leq M $ } \qquad S_{N,\mi}^{\m} := } \\ & & \left\{ \sigma \in S_N^{\m} \quad \begin{array}{|c c} p^{\sigma}_{m_{i_k}} > p^{\sigma}_{m_{i_k + 1}} & \forall\; 1\leq k \leq L \\ p^{\sigma}_{m_j} < p^{\sigma}_{m_j + 1 } & \forall\; j \in \{1,2,\dots,M \}\setminus\{i_1,i_2,\dots ,i_L\} \end{array} \right\} \end{eqnarray*} \end{df} Recall ({\bf 1.4}) that for $ \sigma \in S_N $ we define: \begin{equation*} \{ \sigma_1,\dots,\sigma_N \} := \sigma.\{1,2,\dots,N \} \; ,\quad i = \sigma_{p^{\sigma}_i} \quad (i=1,\dots,N) . \end{equation*} \noindent {\bf Example} Let $N=4\;,\; M=2$ and the motif is: $ (m_1 , m_2) = (1 , 3)$. The corresponding partition is: $ (2,1,1,0) $. In this case the set $ S_N^{(m_1,m_2)} = S_4^{(1,3)} $ contains altogether twelve elements. This set is subdivided into four subsets: $ S_{4,(\emptyset)}^{(1,3)}\;,\; S_{4,(1)}^{(1,3)}\;,\; S_{4,(3)}^{(1,3)}\;,\; S_{4,(1,3)}^{(1,3)} $ : \begin{align*} S_{4,(\emptyset)}^{(1,3)} & = \{ \{ 1234 \} \} \,; \\ S_{4,(1)}^{(1,3)} & = \{ \{ 2134\} ,\{ 2314 \} ,\{ 2341 \} \} \, ; \\ S_{4,(3)}^{(1,3)} & = \{ \{ 1243 \} ,\{ 1423 \} ,\{ 4123 \} \} \, ; \\ S_{4,(1,3)}^{(1,3)} & = \{ \{ 2143\} ,\{ 2413 \} ,\{ 4213 \} ,\{ 2431 \} ,\{ 4231 \} \} . \end{align*} \mbox{} \noindent {\bf 3.} Let us describe several properties of the sets $S_{N,\mi}^{\m} \quad ( (m_1,m_2,\dots,m_M) \in {\frak M}_N \;, \{i_1,i_2,\dots ,i_L\} \subset \{1,2,\dots,M \} \quad (0\leq L\leq M) )$. Throughout this paragraph we fix such a set $S_{N,\mi}^{\m} $. Let $\mu$ be the partition that corresponds to $(m_1,m_2,\dots,m_M)$. \begin{lemma} Let $\sigma \in S_{N,\mi}^{\m} $. then either: \begin{align*} \exists \: i \in \{1,2,\dots,N-1 \} \, , & \; \sigma\prime \in S_{N,\mi}^{\m} \quad \text{{\em s.t.}} \tag{i} \\ & \sigma = (i,i+1)\sigma\prime \: ,\; \mu_{\sigma_i} < \mu_{\sigma_{i+1}} . \\ & \text{{\em and therefore} $ \sigma\prime > \sigma \; ( \Leftrightarrow \mu_{\sigma\prime} > \mu_{\sigma})$}; \end{align*} or \begin{align*} \sigma & = \sigma[0]:= (m_{i_1},m_{i_1}+1)(m_{i_2},m_{i_2}+1)\ldots (m_{i_L},m_{i_L}+1) . \tag{ii} \end{align*} \end{lemma} \begin{pf} \noindent Let $ \sigma = \{\sigma_1,\sigma_2,\dots,\sigma_N\} $. Examine the pairs $ \sigma_i , \sigma_{i+1} \; (i =1,\dots,N-1) $ step by step starting with $i=1$ and increasing $i$ by $1$ at every next step. \noindent At each step $i$ one has the two possibilities: \begin{equation*} {\mathrm I.}\quad \mu_{\sigma_i} \geq \mu_{\sigma_{i+1}} \qquad \qquad {\mathrm II.} \quad \mu_{\sigma_i} < \mu_{\sigma_{i+1}} \end{equation*} \noindent If {\em I.}, go to the next step. If at each step $ (i =1,\dots,N-1)$ holds {\em I.}, then $\sigma = {\mathrm id}$ and therefore $ L=0 \, , \, \sigma[0] = {\mathrm id} $. The poof is finished. \noindent If {\em II.}, then one has the further two possibilities: \begin{equation*} 1. \; (i,i+1)\sigma \not\in S_{N,\mi}^{\m} \qquad \qquad 2. \; (i,i+1)\sigma \in S_{N,\mi}^{\m} . \end{equation*} \noindent If {\em 1.} go to the next step. If for all cases where {\em II.} holds we have {\em 1.}, then $ \sigma = \sigma[0]$. The poof is finished. \noindent If {\em 2.} denote $\sigma\prime := (i,i+1)\sigma $. Since $ \mu_{\sigma_i} < \mu_{\sigma_{i+1}} $, we have $\mu_{\sigma\prime} > \mu_{\sigma} \Leftrightarrow \sigma\prime > \sigma .$ The poof is finished. \end{pf} \noindent The element $\sigma[0]: = (m_{i_1},m_{i_1}+1)(m_{i_2},m_{i_2}+1)\ldots (m_{i_L},m_{i_L}+1) \\ \in S_{N,\mi}^{\m} $ is the maximal element in the set $S_{N,\mi}^{\m}$ (Cf. {\bf 1.4.2} for the definition of the ordering of elements of $S_N^{\m}$). \noindent We summarize the properties of the subset $S_{N,\mi}^{\m} $ in the following proposition: \begin{prop} Let $\{i_1,i_2,\dots ,i_L\} \subset \{1,2,\dots,M \} \;,\; (0\leq L\leq M) $. \noindent Then: \begin{equation*} S_{N,\mi}^{\m} = S_{N,(m_{i_1},\dots,m_{i_L})}^{(m_{i_1},\dots,m_{i_L})}\; ,\tag{i} \end{equation*} \begin{gather*} \forall \; \sigma \in S_{N,\mi}^{\m} \quad \exists \; \{ j_1,j_2,\dots,j_r\} \subset \{ 1,\dots,N-1\} \quad ( r \geq 0 ) \; \text{{\em s.t.}} \tag{ii} \end{gather*} the elements $ \sigma[r] , \sigma[r-1],\dots , \sigma[1] $ defined by \begin{align*} \sigma[0] & := (m_{i_1},m_{i_1}+1)(m_{i_2},m_{i_2}+1)\dots (m_{i_L},m_{i_L}+1), \\ \sigma[k] & := (j_k,j_k+1)\sigma[k-1] \quad ( k=1,2,\dots,r ) \end{align*} belong to the set $ S_{N,\mi}^{\m} $, \\ satisfy \begin{equation*} \sigma[k] < \sigma[k-1] \qquad ( k=1,2,\dots,r ), \end{equation*} and $ \sigma[r] = \sigma $. \\ \mbox{} \noindent {\em (iii)} \\ If there exists $i \in \{1,2,\dots,N\}$ such that \begin{align*} \sigma \; , \; (i,i+1)\sigma \; & \in \; S_{N,\mi}^{\m} , \\ \text{then} & \quad |\sigma_{i+1} - \sigma_i | \geq 2 . \end{align*} If there exists $i \in \{1,2,\dots,N\}$ such that \begin{align*} \sigma \;\in \; S_{N,\mi}^{\m} , & \; (i,i+1)\sigma \; \in \; S_{N,(m_{i_2},\dots,m_{i_L})}^{(m_1,m_2,\dots,m_M)} , \\ \text{then} & \quad \sigma_i \sigma_{i+1} + 1 . \end{align*} \end{prop} \begin{pf} \noindent {\em (i)} is a direct consequence of the {\em Definitions 1 ({\bf 1.4}) and 3}. \noindent {\em (ii)} follows from Lemma 3 and the observation that $\sigma[0] = (m_{i_1},m_{i_1}+1)(m_{i_2},m_{i_2}+1)\ldots (m_{i_L},m_{i_L}+1) $ is the maximal element in $S_{N,\mi}^{\m} $. \noindent {\em (iii)} For $\sigma \in S_{N,\mi}^{\m} $ assume that $ \sigma_{i+1} = \sigma_i +1 $. One has the two possibilities: \begin{align*} 1. & \quad i = p^{\sigma}_j \; , \; i+1 = p^{\sigma}_{j+1} \quad \text{where}\; m_{s-1}+1 \leq j,j+1\leq m_s \\ & \text{for some}\; s \in \{1,2,\dots,M \} \; (m_0 := 0). \\ 2. & \quad i = p^{\sigma}_{m_s}\; , \; i+1 = p^{\sigma}_{m_s+1} \\ & \text{for some} \; s \in \{1,2,\dots,M \} . \end{align*} In the case {\em 1.} $ (i,i+1)\sigma \not\in S_N^{\m} .$ \\ In the case {\em 2.} $ (i,i+1)\sigma \in S_{N,(m_s,m_{i_1},\dots,m_{i_L})}^{(m_1,m_2,\dots,m_M)} . $ \noindent Assume that $ \sigma_i = \sigma_{i+1} + 1 .$ \noindent In this case $ \exists \; s \in \{i_1,i_2,\dots ,i_L\} $ {\em s.t. } \begin{equation*} \sigma_i = m_s + 1 \; , \; \sigma_{i+1} = m_s \quad \Leftrightarrow \quad i = p^{\sigma}_{m_s+1} \; , \;i +1 = p^{\sigma}_{m_s}. \end{equation*} Since by definition of $S_{N,\mi}^{\m}$ : \begin{gather*} i < p_{m_s+2} < \ldots < p_{m_{s+1}} \, , \\ p_{m_{s-1}} < \ldots < p_{m_s-2} < p_{m_s-1}< i+1 ; \end{gather*} we have: \begin{equation*} (i,i+1)\sigma \in S_{N,(m_{i_1},\dots,m_{i_L})\setminus m_s}^{(m_1,m_2,\dots,m_M)}. \end{equation*} The first statement in {\em (iii)} is proven. \noindent The second statement in {\em (iii)} is proven in a similar way. \end{pf} \subsection{A property of the eigenvectors $\varphi_{\sigma}^{\l}$ for $\l \in {\frak M}_N$} {\bf 1.} In this section we shall derive a certain proprerty of the eigenspaces ${\cal E}^{\mu}\quad (\mu \in {\frak M}_N)$ of the operators $\Delta_1(u) , y_i (i=1,\dots,N). $ First of all, we notice that the eigenvalue $ \delta^{\mu}_1(u) $ of $ \Delta_1(u) $ associated with ${\cal E}^{\mu}$ can be represented in the additive ``particle'' form: \begin{equation} \delta^{\mu}_1(u) := \delta^{(m_1,m_2,\dots,m_M)}_1(u) = \sum_{k=1}^M \delta^{(m_k)}_1(u), \end{equation} where $ \delta^{(m)}_1(u) \quad m \in \{1,\dots,N-1\}$ is the one-particle eigenvalue: \begin{equation} \delta^{(m)}_1(u) := u\sum_{i=1}^m \prod\begin{Sb} 1\leq j\leq N \\ j\neq i \end{Sb} (1 + u q^{l_j})q^{l_i} . \end{equation} (Cf. 2.2.26 for the definition of $ l_i $ ). \noindent {\bf 2.} Let $ \mu = (m_1,m_2,\dots,m_M) \in {\frak M}_N$ then the conjugate partition $\mu\prime$ is: $\mu\prime = (m_M,m_{M-1},\dots,m_1)$. Due to (2.2.9,.10) we have: \begin{equation} \Delta_1(u) e_{m_1}e_{m_2}\dots e_{m_M} = \left(\sum_{k=1}^M \delta^{(m_k)}_1(u)\right)e_{m_1}e_{m_2}\dots e_{m_M}. \end{equation} While for any symmetric polynomial $S$ one has: \begin{equation} y_i S = q^{l_i} S \quad (i=1,\dots,N) . \end{equation} \mbox{} \noindent Now we have the following proposition: \begin{prop} Let $ (m_1,m_2,\dots,m_M) \in {\frak M}_N\:,\; \{i_1,i_2,\dots ,i_L\} \subset \{1,2,\dots,M \} \;\\ (0\leq L\leq M), $ and $ \{j_1,\dots,j_{M-L}\} = \{1,2,\dots,M \}\setminus\{i_1,i_2,\dots ,i_L\}. $ \noindent Then: \begin{equation} \varphi^{(m_1,m_2,\dots,m_M)}_{\sigma} \; (\sigma \in S_{N,\mi}^{\m})\; = \varphi^{(m_{i_1},\dots,m_{i_L})}_{\sigma} e_{m_{j_1}}e_{m_{j_2}}\dots e_{m_{j_{M-L}}}. \end{equation} \end{prop} \begin{pf} Compute the action of the operators $\Delta_1(u) , y_i \; (i=1,\dots,N). $ on the polynomial $ \varphi^{(m_{i_1},\dots,m_{i_L})}_{\sigma} e_{m_{j_1}}e_{m_{j_2}}\dots e_{m_{j_{M-L}}}.$ \begin{align} y_i\varphi^{(m_{i_1},\dots,m_{i_L})}_{\sigma}S & = q^{l_{\sigma_i}}\varphi^{(m_{i_1},\dots,m_{i_L})}_{\sigma}S \quad (i=1,\dots,N). \tag{a} \end{align} For any symmetric polynomial $ S $. \begin{multline} \Delta_1(u)\varphi^{(m_{i_1},\dots,m_{i_L})}_{\sigma}e_{m_{j_1}}e_{m_{j_2}}\dots e_{m_{j_{M-L}}} = \\ = [\Delta_1(u)\,,\,\varphi^{(m_{i_1},\dots,m_{i_L})}]e_{m_{j_1}}e_{m_{j_2}}\dots e_{m_{j_{M-L}}} + \tag{b} \\ + \varphi^{(m_{i_1},\dots,m_{i_L})}\Delta_1(u) e_{m_{j_1}}e_{m_{j_2}}\dots e_{m_{j_{M-L}}}. \end{multline} Due to the special form of $\Delta_1(u)$ (Cf. {\em Proposition 4}) the commutator in the last formula is a zero-order differential operator. Therefore if $ S $ is a symmetric polynomial we have: \begin{equation} [\Delta_1(u)\,,\,\varphi^{(m_{i_1},\dots,m_{i_L})}]S = f(u)S, \end{equation} where $f(u)$ is a function of $ z_i \; (i=1,\dots,N) $ independent of $S$. Furthermore: \begin{equation*} \Delta_1(u) 1 = 0, \end{equation*} and therefore: \begin{align*} [\Delta_1(u)\,,\,\varphi^{(m_{i_1},\dots,m_{i_L})}]1 & =\Delta_1(u) \varphi^{(m_{i_1},\dots,m_{i_L})} = \\ & = \sum_{1\leq k\leq L}^M \delta^{(m_{i_k})}_1(u)\varphi^{(m_{i_1},\dots,m_{i_L})}. \end{align*} Taking {\em (b),(4.4.7,.11)} into account we get: \begin{multline*} \Delta_1(u)\varphi^{(m_{i_1},\dots ,m_{i_L})}_{\sigma}e_{m_{j_1}}e_{m_{j_2}}\dots e_{m_{j_{M-L}}} = \\ = \left(\sum_{1\leq k\leq L} \delta^{(m_{i_k})} + \sum_{1\leq s\leq M - L} \delta^{(m_{j_s})} \right) \varphi^{(m_{i_1},\dots ,m_{i_L})}_{\sigma}e_{m_{j_1}}e_{m_{j_2}}\dots e_{m_{j_{M-L}}}= \\ = \sum_{1\leq i\leq M} \delta^{(m_i)} \varphi^{(m_{i_1},\dots ,m_{i_L})}_{\sigma}e_{m_{j_1}}e_{m_{j_2}}\dots e_{m_{j_{M-L}}}. \end{multline*} \noindent Since the joint spectrum of $ \Delta_1(u) \,,\,y_i \; (i=1,\dots,N) $ is multiplicity-free, we conclude from {\em (a)} and the last equation that: \begin{equation} \varphi^{(m_{i_1},\dots,m_{i_L})}_{\sigma}e_{m_{j_1}}e_{m_{j_2}}\dots e_{m_{j_{M-L}}} = {\mathrm {const}}\, \varphi^{(m_1,m_2,\dots,m_M)}_{\sigma} \quad ( \sigma \in S_{N,\mi}^{\m} ) . \end{equation} By comparison of the maximal monomials in the both sides of the equation {\em const} $ =1$. \end{pf} \mbox{} \noindent Since $ e_r ( z_1=\omega^1,\dots,z_N = \omega^N ) = 0 \quad (1\leq r \leq N-1)$ we have the following corollary to Proposition 9: \begin{cor} Let $ (m_1,m_2,\dots,m_M) \in {\frak M}_N\:,\; M > 0 \:,\; \sigma \in S_N^{\m} $. \noindent Then: \begin{align*} \varphi^{(m_1,m_2,\dots,m_M)}_{\sigma}( z_1=\omega^1,\dots,z_N = \omega^N )& = 0 \\ \text{{\em unless}} & \quad \sigma \in S^{(m_1,m_2,\dots,m_M)}_{N,(m_1,m_2,\dots,m_M)}. \end{align*} \end{cor} \subsection{Limit $q\rightarrow 0$ of the eigenfunctions of $\Delta_1(u) , y_i\quad (i=1,\dots,N)$} {\bf 1.} The aim of this subsection is to compute some of the eigenfunctions $\varphi_{\sigma}^{\l}\quad (\l\in {\frak M}_N,\sigma \in \S)$ in the limit $q\rightarrow 0$ {\em under assumption} that this limit is well-defined. We do not have a proof of the last statement. In particular examples where $N$ is small this statement holds. \mbox{} \noindent Introduce operators: \begin{equation} \gamma_{i,j} := qg_{i,j}|_{q=0} = \frac{z_i}{z_i -z_j}(K_{i,j}- 1) \quad (i\neq j\in \{1,\dots,N\} ). \end{equation} \begin{lemma} Let $(m_1,m_2,\dots,m_M) \in {\frak M}_N$, and $\psi\in{\Bbb C}[z_1,\dots,z_N]$ satisfies the equations: \begin{align} (\gamma_{m_i,m_i+1} + 1 )\psi & = 0 \quad ( i=1,\dots,M), \\ \gamma_{n,n+1}\psi & = 0 \quad ( n\in \{1,\dots,N-1\}\setminus\{m_1,\dots,m_M\}). \end{align} Then \begin{equation} \psi = (z_1\dots z_{m_1})^M(z_{m_1+1}\dots z_{m_2})^{M-1}\ldots(z_{m_{M-1}+1}\dots z_{m_M}) S(z_1,\dots,z_N) , \end{equation} where $S$ is a symmetric polynomial. \end{lemma} \begin{pf} Let $N=2$ , $\psi = \psi(z_1,z_2)$. The equation $ (\gamma_{1,2}+ 1)\psi = 0 $ implies that \begin{equation} \frac{z_1\psi(z_2,z_1) - z_2\psi(z_1,z_2)}{z_1 - z_2} = 0. \end{equation} This leads to $ \psi(z_1,z_2) = z_1 S(z_1,z_2)$, where $S$ is a symmetric polynomial. The equation \begin{equation} \gamma_{1,2}\psi = \frac{z_1(\psi(z_2,z_1) - \psi_(z_1,z_2))}{z_1 - z_2} = 0 , \end{equation} yields $ \psi(z_1,z_2) = S(z_1,z_2) $ where $ S$ is a symmetric polynomial. The case of arbitrary $N$ reduces to the case $ N=2$ by consideration of consequtive pairs of coordinates. \end{pf} \mbox{} \noindent {\bf 2.} We {\em conjecture} that the limit $q \rightarrow 0$ of $\varphi_{\sigma}^{\l}$ is well-defined: \begin{equation} \varphi_{\sigma}^{\l}|_{q=0}:= \varphi_{\sigma}^{\l,0} = z^{\l_{\sigma}} + \text{``s.m.''} \qquad (\l \in {\frak M}_N ,\; \sigma\in \S). \end{equation} In what follows we assume that the statement of the conjecture is valid. Fix $(m_1,m_2,\dots,m_M) \in {\frak M}_N $. Our purpose is to find $ \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0}$ for $ \sigma[0]:= (m_1,m_1+1)\dots(m_M,m_M+1).$ Take the limit $q\rightarrow 0$ in the eq. (4.4.4) of Proposition 9. This yields \begin{align} (\gamma_{m_i,m_i+1} + 1) \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0} & = \varphi_{(m_i,m_i+1)\sigma[0]}^{(m_1,m_2,\dots,m_M),0}\quad (i = 1,\dots,M),\\ \gamma_{n,n+1} \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0} & = 0 \quad (n \in \{1,\dots,N-1\}\setminus\{m_1,\dots,m_M\}). \end{align} According to Proposition 12, we have \begin{multline} \varphi_{(m_i,m_i+1)\sigma[0]}^{(m_1,m_2,\dots,m_M),0} = \varphi_{(m_1,m_1+1)\dots\widehat{(m_i,m_i+1)}\dots (m_M,m_M+1)}^{(m_1,m_2,\dots,m_M),0} = \\ \varphi_{(m_1,m_1+1)\dots\widehat{(m_i,m_i+1)}\dots (m_M,m_M+1)}^{(m_1,\dots,\hat{m_i},\dots,m_M ),0} e_{m_i} \quad (i =1,\dots,M). \end{multline} Where we put a hat over terms that are omitted. The pair of equations (4.4.20,.21) provides a set of recurrent relations for the eigenfunctions $\varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0}$. Notice that when $M=0$ we have $ \varphi_{\sigma[0]}^{(\emptyset),0} = 1$. Taking into account Lemma 4 we write the general solution of these recurrent relations: \begin{align} \varphi_{\sigma[0]}^{(\emptyset),0} & = 1, \\ \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0} & = (e_{m_1} - z_1\dots z_{m_1})(e_{m_2} - z_1\dots z_{m_2})\ldots (e_{m_M} - z_1\dots z_{m_M})+ \notag\\ & + (z_1\dots z_{m_1})^M(z_{m_1+1}\dots z_{m_2})^{M-1}\ldots(z_{m_{M-1}+1}\dots z_{m_M})\times \notag \\ & \times S^{(m_1,m_2,\dots,m_M)}(z_1,\dots,z_N) \quad ( M \geq 1). \notag \end{align} Where $ S^{(m_1,m_2,\dots,m_M)}(z_1,\dots,z_N) $ is an arbitrary symmetric polynomial. Observe now that the total degree of the homogeneous polynomial $ \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0} $ is equal to $ m_1 + \dots + m_M \quad ((m_1,m_2,\dots,m_M) \in {\frak M}_N) $. Therefore we must have $ S^{(m_1,m_2,\dots,m_M)}(z_1,\dots,z_N) = {\mathrm {const}} $. Observe further that \begin{multline} (z_1\dots z_{m_1})^M(z_{m_1+1}\dots z_{m_2})^{M-1}\ldots(z_{m_{M-1}+1}\dots z_{m_M}) > \\ > {\mathrm {max}}( \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0} ) \quad ( \sigma[0]:= (m_1,m_1+1)\dots(m_M,m_M+1) ). \end{multline} Where $ {\mathrm {max}}( \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0} ) $ is the maximal monomial of $ \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0}. $ Therefore we must have: $ S^{(m_1,m_2,\dots,m_M)}(z_1,\dots,z_N) = 0 $. Thus for $ (m_1,m_2,\dots,m_M) \in {\frak M}_N $ : \begin{multline} \varphi_{\sigma[0]}^{(\emptyset),0} = 1, \\ \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0} = (e_{m_1} - z_1\dots z_{m_1})(e_{m_2} - z_1\dots z_{m_2})\ldots (e_{m_M} - z_1\dots z_{m_M})\notag \\ ( M \geq 1 ) \notag \end{multline} Notice that \begin{multline} \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M),0}(z_1=\omega^1,\dots,z_N=\omega^N ) = (-1)^M \omega^{\frac{1}{2}\sum_{i=1}^M m_i(m_i+1)} , \\ \quad ( (m_1,m_2,\dots,m_M) \in {\frak M}_N , \; \sigma[0]= (m_1,m_1+1)\dots(m_M,m_M+1)). \end{multline} \section{Hecke-invariant (``bosonic'') subspaces of ${\cal E}^{\l}\otimes H$ for ${\l} \in {\frak M}_N $} \subsection{Preliminaries} {\bf 1.} Let ${\cal E}^{\l} \subset {\Bbb C}[z_1,\dots,z_N]$ be the eigenspace of the operators $ \Delta_1(u)\,,\,y_i\quad(i=1,\dots,N)$ parametrized by a partition $\l$ and $ H := ({\Bbb C}^2)^{\otimes N}$. The bosonic subspace ${\cal B}^{\l}$ of ${\cal E}^{\l}\otimes H $ is defined as follows: \begin{equation} B^{\l}:= \{ b \in {\cal E}^{\l}\otimes H | (g_{i,i+1} -t_{i,i+1})b = 0\quad (i=1,\dots,N-1) \}. \end{equation} Since $ {\cal P}:={\Bbb C}[z_1,\dots,z_N]\otimes H = \oplus_{\l}({\cal E}^{\l}\otimes H)$ and $ g_{i,i+1} :{\cal E}^{\l}\mapsto {\cal E}^{\l}\quad(i=1,\dots,N-1) ;$ we have: $ {\cal B} =\oplus_{\l}{\cal B}^{\l}. $ (Cf. (1.1.36), (3.3.7) for the definition of ${\cal B}$). \noindent {\bf 2.} Any vector $\psi$ from ${\cal E}^{\l}\otimes H$ is represented as follows: \begin{equation} \psi = \sum_{\sigma\in\S}\varphi_{\sigma}^{\l} \chi_{\sigma} , \end{equation} where $\chi_{\sigma} \quad (\sigma \in \S) \in H $. \mbox{} \noindent The condition $ (g_{i,i+1} -t_{i,i+1})\psi = 0\quad(i=1,\dots,N-1) $ gives a set of linear equations which must be satisfied by the vectors $\chi_{\sigma} \quad (\sigma \in \S) \in H .$ In order to derive these equations we can apply the result of Proposition 9 to find out the action of $g_{i,i+1}\quad(i=1,\dots,N-1) $ on $\psi$, and then use the linear-independence of the polynomials $ \varphi_{\sigma}^{\l}\quad (\sigma\in\S)$. In this way we arrive at the following proposition: \begin{prop} A vector $ \psi \in {\cal E}^{\l}\otimes H ;\; \psi = \sum_{\sigma\in\S}\varphi_{\sigma}^{\l} \chi_{\sigma} $ belongs to $ {\cal B}^{\l} $ iff $\chi_{\sigma} \quad (\sigma \in \S) \in H $ satisfy the following set of equations $ (i=1,\dots,N-1) $: \begin{gather*} q\frac{(q^{-1}q^{l_{\sigma_{i+1}}} - qq^{l_{\sigma_i}})(q^{-1}q^{l_{\sigma_i}} -qq^{l_{\sigma_{i+1}}})}{q^{l_{\sigma_i}} - q^{l_{\sigma_{i+1}}}}\chi_{(i,i+1)\sigma} =\tag{a} \\ = \left( (q^{l_{\sigma_{i+1}}}-q^{l_{\sigma_i}})t_{i,i+1} - \dq^{l_{\sigma_{i+1}}}\right)\chi_{\sigma} \\ \text{{\em when}}\quad \l_{\sigma_{i+1}} > \l_{\sigma_i }\;, \\ (t_{i,i+1} - q ) \chi_{\sigma} = 0 \tag{b} \\ \text{{\em when}}\quad \l_{\sigma_{i+1}} = \l_{\sigma_i } \Leftrightarrow (i,i+1)\sigma \not\in \S \Rightarrow \sigma_{i+1} = \sigma_i+1 \; , \\ \chi_{(i,i+1)\sigma} = \frac{ (q^{l_{\sigma_{i+1}}} -q^{l_{\sigma_i}})t_{i,i+1} - \dq^{l_{\sigma_{i+1}}}}{q^{-1}(q^{l_{\sigma_{i+1}}} - q^{l_{\sigma_i}})}\chi_{\sigma} \tag{c}\\ \text{{\em when}}\quad \l_{\sigma_{i+1}} < \l_{\sigma_i } \; . \end{gather*} \end{prop} \subsection{ Spaces ${\cal B}^{\mu}$ for $ \mu \in {\frak M}_N $ } {\bf 1.} Let us fix a partition $\mu \in {\frak M}_N$ parametrized by a motif $ (m_1,m_2,\dots,m_M) $. We use the same notation $(m_1,m_2,\dots,m_M)$ for both the motf and the partition. \noindent For $\mu \in {\frak M}_N$ let us further analyse the equations (a)-(c) obtained in the Proposition 10. In section {\bf 4.3} (Definition 3) we introduced the decomposition of the set $ S_N^{\m} $ into disjoint subsets: \begin{equation} S_N^{\m} = \bigsqcup_{\{i_1,i_2,\dots ,i_L\} \in \{1,2,\dots,M \}} S_{N,\mi}^{\m} \qquad (0\leq L\leq M) \end{equation} This decomposition is reflected in the equations (a)-(c) of the Proposition 10. \noindent {\bf 2.} Let in these equations $ \sigma\,,\, i $ be such that $ \sigma \, ,\, (i,i+1)\sigma \in S_{N,\mi}^{\m} $ for some $ \{i_1,i_2,\dots ,i_L\} \in \{1,2,\dots,M \} $. According to Proposition 8,(iii) $ |\sigma_i - \sigma_{i+1}| \geq 2.$ Let $ \sigma_i - \sigma_{i+1} \geq 2 $ and consequently $ \mu_{\sigma_i} < \mu_{\sigma_{i+1}} $. Since $ \sigma_i - \sigma_{i+1} \geq 2 $ the coefficient in front of $ \chi_{(i,i+1)\sigma} $ in Pr.10(a) is not equal to zero. Therefore we have: \begin{multline} \chi_{(i,i+1)\sigma} = \frac{(q^{l_{\sigma_i}}-q^{l_{\sigma_{i+1}}})((q^{l_{\sigma_{i+1}}}-q^{l_{\sigma_i}})t_{i,i+1} -\dq^{l_{\sigma_{i+1}}})}{q(q^{-1}q^{l_{\sigma_{i+1}}}-qq^{l_{\sigma_i}})(q^{-1}q^{l_{\sigma_i}}-qq^{l_{\sigma_{i+1}}})}\chi_{\sigma} \\ ( \sigma , (i,i+1)\sigma \in S_{N,\mi}^{\m}\, ,\, \sigma_i - \sigma_{i+1} \geq 2 ). \end{multline} \mbox{} \noindent Let $\sigma_i - \sigma_{i+1} \leq -2$ and consequently $ \mu_{\sigma_i} > \mu_{\sigma_{i+1}} $. Pr.10(c) gives: \begin{multline} \chi_{(i,i+1)\sigma} = \frac{((q^{l_{\sigma_{i+1}}}-q^{l_{\sigma_i}})t_{i,i+1} -\dq^{l_{\sigma_{i+1}}})}{q^{-1}(q^{l_{\sigma_{i+1}}}-q^{l_{\sigma_i}})}\chi_{\sigma} \\ ( \sigma , (i,i+1)\sigma \in S_{N,\mi}^{\m}\, ,\, \sigma_i - \sigma_{i+1} \leq -2 ). \end{multline} \mbox{} \noindent Introduce a pair of $U$-intertwiners: \begin{equation} Y^{\pm}(z) : = \varrho^{\pm}(z)\frac{zt - t^{-1}}{q^{-1}z - q} \; \in End(V\otimes V), \end{equation} where \begin{equation} \varrho^+(z) := \frac{z - 1}{q^2z - 1 }\; , \qquad \varrho^-(z) := \frac{z - q^2}{z - 1 }. \end{equation} The eq. (5.5.4,.5) can be reformulated as follows: Let $ \sigma , (i,i+1)\sigma \in S_{N,\mi}^{\m} \\ \quad (\{i_1,i_2,\dots ,i_L\} \in \{1,2,\dots,M \} \quad (0\leq L\leq M) ) $ then: \begin{equation} \chi_{(i,i+1)\sigma} = \begin{cases} Y^+_{i,i+1}(q^{l_{\sigma_i}-l_{\sigma_{i+1}}}) \chi_{\sigma} & \text{when $ \sigma_i - \sigma_{i+1} \geq 2 \Rightarrow \mu_{\sigma_i} < \mu_{\sigma_{i+1}}$ }, \\ Y^-_{i,i+1}(q^{l_{\sigma_i}-l_{\sigma_{i+1}}}) \chi_{\sigma} & \text{when $ \sigma_i - \sigma_{i+1} \leq -2 \Rightarrow \mu_{\sigma_i} > \mu_{\sigma_{i+1}}$}. \end{cases} \end{equation} Notice that when $ \sigma , (i,i+1)\sigma \in S_{N,\mi}^{\m} $ the intertwiners $Y^{\pm}_{i,i+1}(q^{l_{\sigma_i}-l_{\sigma_{i+1}}})$ are invertible. \mbox{} \noindent {\bf 3.} Now consider the situation when $ \sigma$ and $(i,i+1)\sigma$ in Proposition 10 belong to different subsets $S_{N,\mi}^{\m}$. According to Proposition 8 this situation takes place when $|\sigma_i - \sigma_{i+1} | = 1.$ \noindent Let $ \sigma_{i+1}= \sigma_i + 1$ and consequently $ \mu_{\sigma_i} > \mu_{\sigma_{i+1}}.$ In this case $ \exists \; s \in \{1,2,\dots,M \} $ s.t. $ \sigma_i = m_s \, , \: \sigma_{i+1} = m_s + 1 $ or , equivalently , $ i = p^{\sigma}_{m_s} \,,\: i+1 = p^{\sigma}_{m_s+1}$; and if $ \sigma \in S_{N,\mi}^{\m} $ then $ s \not\in \{i_1,i_2,\dots ,i_L\} .$ Since $ ((i,i+1)\sigma)_i = m_s+1\;,\; ((i,i+1)\sigma)_{i+1} = m_s ;$ we have $ (i,i+1)\sigma \in S_{N,(m_s,m_{i_1},\dots,m_{i_L})}^{(m_1,m_2,\dots,m_M)} $. \noindent Substituting $ \sigma_{i+1}= \sigma_i + 1$ into Pr.10(c) we find: \begin{multline} \chi_{(i,i+1)\sigma} = q(t_{i,i+1} - q)\chi_{\sigma} = -(q^2 + 1)\Pi_{i,i+1}^-(q)\chi_{\sigma} \\ ( \sigma \in S_{N,\mi}^{\m} \;,\; (i,i+1)\sigma \in S_{N,(m_s,m_{i_1},\dots,m_{i_L})}^{(m_1,m_2,\dots,m_M)}\;,\; \\ \sigma_i = m_s , \sigma_{i+1} = m_s + 1 ( s \in \{1,2,\dots,M \}\setminus\{i_1,i_2,\dots ,i_L\} ) ). \end{multline} (Cf. (1.1.12,.13) for the definition of projectors $\Pi_{i,i+1}^{\pm}(q)$). \mbox{} \noindent Let $ \sigma_i= \sigma_{i+1} + 1$ and consequently $ \mu_{\sigma_i} < \mu_{\sigma_{i+1}}.$ In this case $ \exists \; s \in \{1,2,\dots,M \} $ s.t. $ \sigma_i = m_s+1 \, , \: \sigma_{i+1} = m_s $ ; and since $ \sigma \in S_{N,\mi}^{\m} \: ; \; s \in \{i_1,i_2,\dots ,i_L\} .$ On the other hand $ (i,i+1)\sigma \in S_{N,(m_{i_1},\dots,m_{i_L})\setminus m_s}^{(m_1,m_2,\dots,m_M)}.$ \\ Here $(m_{i_1},\dots,m_{i_L})\setminus m_s$ signifies the motif obtained from $( m_{i_1},\dots,m_{i_L})$ by removing $m_s$. \noindent Substituting $ \sigma_i= \sigma_{i+1} + 1$ into Pr.10(a) we find: \begin{multline} (t_{i,i+1} + q^{-1})\chi_{\sigma} = 0 \quad \Rightarrow \quad \Pi_{i,i+1}^+(q)\chi_{\sigma} = 0 \\ ( \sigma \in S_{N,\mi}^{\m} \;,\; (i,i+1)\sigma \in S_{N,(m_{i_1},\dots,m_{i_L})\setminus m_s}^{(m_1,m_2,\dots,m_M)}\;,\; \\ \sigma_i = m_s+1 , \sigma_{i+1} = m_s ( s \in \{i_1,i_2,\dots ,i_L\} \subset \{1,2,\dots,M \} ) ). \end{multline} \mbox{} \noindent {\bf 4.} Now we are in position to reformulate Propositon 10 in a more suitable form in the case when $ \l \equiv \mu \equiv (m_1,m_2,\dots,m_M) \in {\frak M}_N $. \begin{prop} Let $ \psi \in {\cal E}^{\mu}\otimes H$, i.e. \begin{equation} \psi = \sum_{\sigma \in S_N^{\m}}\varphi^{(m_1,m_2,\dots,m_M)}_{\sigma} \chi_{\sigma} \qquad ( \chi_{\sigma} \in H ). \end{equation} Then $ \psi \in {\cal B}^{(m_1,m_2,\dots,m_M)} $ iff $ \chi_{\sigma} \quad (\sigma \in S_N^{\m}) $ satisfy the following linear relations: \begin{multline} \forall \; \{i_1,i_2,\dots ,i_L\} \subset \{1,2,\dots,M \} \quad (0\leq L\leq M) \\ \chi_{(m_{i_1},m_{i_1}+1)\dots (m_{i_L},m_{i_L}+1)} = \\ = -(q^2 + 1) \Pi_{m_{i_k},m_{i_k}+1}^-(q) \chi_{(m_{i_1},m_{i_1}+1)\dots \widehat{ (m_{i_k},m_{i_k}+1)}\dots (m_{i_L},m_{i_L}+1)} \\ (k = 1,2,\dots,L). \end{multline} Where $\widehat{\mbox{}}$ means that the corresponding factor is omitted from the product. \begin{multline} \forall \; \{i_1,i_2,\dots ,i_L\} \subset \{1,2,\dots,M \} \quad (0\leq L\leq M) \\ \text{{\em and }} \; j \in \{1,\dots,N-1\} \; \text{{\em s.t.}} \\ \{j,j+1\}\cap \{m_1,m_1+1,m_2,m_2+1,\dots,m_M,m_M+1\} = \emptyset \; ; \\ \Pi_{j,j+1}^-(q)\chi_{(m_{i_1},m_{i_1}+1)\dots (m_{i_L},m_{i_L}+1)} = 0. \end{multline} \begin{gather} \forall \; \sigma \in S_{N,\mi}^{\m} \quad ( \{i_1,i_2,\dots ,i_L\} \subset \{1,2,\dots,M \} \quad (0\leq L\leq M) ) \\ \chi_{\sigma} = {\Bbb Y}(\sigma)\chi_{(m_{i_1},m_{i_1}+1)\dots (m_{i_L},m_{i_L}+1)}. \notag \end{gather} Where invertible ${\Bbb Y}(\sigma)\;\in\; End(H)$ is recursively defined as follows: \begin{gather*} {\Bbb Y}((m_{i_1},m_{i_1}+1)\dots (m_{i_L},m_{i_L}+1)) := {\mathrm {Id}} \, , \\ \text{{\em for}} \; (i,i+1)\sigma \, \in \, S_{N,\mi}^{\m} \\ {\Bbb Y}((i,i+1)\sigma) = \begin{cases} Y^+_{i,i+1}(q^{l_{\sigma_i}-l_{\sigma_{i+1}}}){\Bbb Y}(\sigma) & \text{{\em if}}\; \sigma_i - \sigma_{i+1} \geq \mbox{} 2\, , \\ Y^-_{i,i+1}(q^{l_{\sigma_i}-l_{\sigma_{i+1}}}){\Bbb Y}(\sigma) & \text{{\em if}}\; \sigma_i - \sigma_{i+1} \leq -2\, . \end{cases} \end{gather*} \end{prop} \mbox{} \noindent It is possible to give more explicit expression for the matrix $ {\Bbb Y}(\sigma) $ that appears in (5.5.14). In notation of Proposition 8 (ii) we have: \begin{align} \text{For} & \; \sigma \in S_{N,\mi}^{\m} \qquad ( \{i_1,i_2,\dots ,i_L\} \subset \{1,2,\dots,M \} \; (0\leq L\leq M) ) \notag \\ {\Bbb Y}(\sigma) & = Y^-_{j_r,j_r+1}(q^{l_{\sigma [r-1]_{j_r}}-l_{\sigma [r-1]_{j_r+1}}}) \dots Y^-_{j_1,j_1+1}(q^{l_{\sigma [0]_{j_1}}-l_{\sigma [0]_{j_1+1}}}). \end{align} Where: \begin{align} \sigma&[k] := (j_k,j_k+1)\dots (j_1,j_1+1)\sigma[0] \in S_{N,\mi}^{\m} \; ( 1\leq k \leq r), \notag \\ \sigma&[0] := (m_{i_1},m_{i_1}+1)\dots (m_{i_L},m_{i_L}+1) , \notag \\ \sigma&[r] = \sigma \notag \\ \sigma&[k-1]_{j_k} - \sigma[k-1]_{j_k+1} \leq -2 \quad (1 \leq k \leq r ). \end{align} \mbox{} \noindent {\bf 5.} Introduce two definitions.\\ For $ (m_1,m_2,\dots,m_M) \in {\frak M}_N $ : \begin{equation} {\cal Z}^{(m_1,m_2,\dots,m_M)}:= \left\{ v \in H \; \begin{array}{|c } \Pi_{j,j+1}^-(q)v = 0 \quad \text{for all}\quad j\in\{1,\dots,N-1\}\;\text{s.t.} \\ \{j,j+1\}\cap\{m_1,m_1+1,\dots,m_M,m_M+1\}=\emptyset \end{array} \right\} \end{equation} More explicitely: \begin{multline*} {\cal Z}^{(m_1,m_2,\dots,m_M)} = S_q(\underset{1}{V}\otimes\dots\otimes\underset{m_1-1}{V})\otimes\underset{m_1}{V}\otimes\underset{m_1+1}{V}\otimes S_q(\underset{m_1+2}{V}\otimes\dots\otimes\underset{m_2-1}{V})\otimes\underset{m_2}{V}\otimes\underset{m_2+1}{V}\otimes\ \\ \ldots \otimes S_q(\underset{m_M+2}{V}\otimes\dots\otimes\underset{N}{V}) \; \subset \; H := \underset{1}{V}\otimes\underset{2}{V}\otimes\ldots\otimes\underset{N}{V}. \end{multline*} Where $S_q$ means $q$-symmetrization and the subscripts indicate positions of the factors in the tensor product $ V^{\otimes N}.$ \\ \mbox{} \\ For any $ \{i_1,i_2,\dots ,i_L\} \in \{1,2,\dots,M \} \quad (0\leq L\leq M) $ define the following projector: \begin{equation} \Pi^{-,(m_1,m_2,\dots,m_M)}_{(m_{i_1},\dots,m_{i_L})}:= \begin{cases}\Pi^-_{m_{i_1},m_{i_1}+1}(q)\dots\Pi^-_{m_{i_L},m_{i_L}+1}(q) & \quad (1\leq L \leq M), \\ {\mathrm I} & (L =0) .\end{cases} \end{equation} \mbox{} \noindent {\bf 6.} Proposition 11 yields the following expression for $ {\cal B}^{\mu}\: \equiv \:{\cal B}^{(m_1,m_2,\dots,m_M)} $: \begin{equation} {\cal B}^{(m_1,m_2,\dots,m_M)} = \BU^{(m_1,m_2,\dots,m_M)}(z){\cal Z}^{(m_1,m_2,\dots,m_M)}. \end{equation} Where \begin{multline} \BU^{(m_1,m_2,\dots,m_M)}(z) := \\ \sum\begin{Sb} I \subset \{1,2,\dots,M \} \\ I:=\{i_1,i_2,\dots ,i_L\} \end{Sb}(-(q^2+1))^L\left\{\sum_{\sigma\inS_{N,\mi}^{\m}}\varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}(z){\Bbb Y}(\sigma)\right\}\Pi^{-,(m_1,m_2,\dots,m_M)}_{(m_{i_1},\dots,m_{i_L})} \; , \\ \BU^{(m_1,m_2,\dots,m_M)}(z) : H \mapsto {\Bbb C}[z_1,\dots,z_N]\otimes H . \end{multline} In the last formula we explicitely indicated $z$-dependence of the polynomials $\varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}$. \mbox{} \noindent {\bf 7.} The space $ {\cal B}^{(m_1,m_2,\dots,m_M)}$ is a $U$-module where the action of $U$ is given by (2.2.19): \begin{equation} T_a^0(u) = L_a(u;\{y_i\}):= L_{a1}(uy_1)L_{a2}(uy_2)\dots L_{aN}(uy_N) \; \in \; End({\Bbb C}[z_1,\dots,z_N]\otimes H). \end{equation} The space ${\cal Z}^{(m_1,m_2,\dots,m_M)}$ is a (reducible,indecomposable) $U$-module as well, with the $U$-action: \begin{equation} L_a(u;\{q^{l_i}\}):= L_{a1}(uq^{l_1})L_{a2}(uq^{l_2})\dots L_{aN}(uq^{l_N}) \; \in \; End(H). \end{equation} The operator $\BU^{(m_1,m_2,\dots,m_M)}(z)$ in (5.5.19,.20) is an $U$-intertwiner of these two modules. To see this let us consider the product $T_a^0(u)\BU^{(m_1,m_2,\dots,m_M)}(z)$.Since $\varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}(z)$ are eigenvectors of $ y_i \; (i=1,\dots,N) $ (Cf. Proposition 6) we get: \begin{multline} T_a^0(u)\BU^{(m_1,m_2,\dots,m_M)}(z) = \\ \sum\begin{Sb} I \subset \{1,2,\dots,M \} \\ I:=\{i_1,i_2,\dots ,i_L\} \end{Sb}(-(q^2+1))^L \; \times \\ \times \; \left\{\sum_{\sigma\inS_{N,\mi}^{\m}}\varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}(z) L_a(u;\{q^{l_{\sigma_i}}\}){\Bbb Y}(\sigma)\right\}\Pi^{-,(m_1,m_2,\dots,m_M)}_{(m_{i_1},\dots,m_{i_L})}. \end{multline} It follows from the recursive definition of ${\Bbb Y}(\sigma)\quad (\sigma \in S_{N,\mi}^{\m})$ given in Proposition 11 that ${\Bbb Y}(\sigma)$ is an $U$-intertwiner: \begin{equation} L_a(u;\{q^{l_{\sigma_i}}\}){\Bbb Y}(\sigma) = {\Bbb Y}(\sigma)L_a(u;\{q^{l_{\sigma[0]_i}}\}). \end{equation} Where $ \sigma[0] = (m_{i_1},m_{i_1}+1)\dots (m_{i_L},m_{i_L}+1)$.\\ Furthermore the projector $\Pi^{-,(m_1,m_2,\dots,m_M)}_{(m_{i_1},\dots,m_{i_L})}$ is an interwiner as well (Cf. 1.1.28): \begin{equation} L_a(u;\{q^{l_{\sigma[0]_i}}\}\Pi^{-,(m_1,m_2,\dots,m_M)}_{(m_{i_1},\dots,m_{i_L})} = \Pi^{-,(m_1,m_2,\dots,m_M)}_{(m_{i_1},\dots,m_{i_L})}L_a(u;\{q^{l_i}\}). \end{equation} Hence we obtain: \begin{equation} T_a^0(u)\BU^{(m_1,m_2,\dots,m_M)}(z) = \BU^{(m_1,m_2,\dots,m_M)}(z)L_a(u;\{q^{l_i}\}). \end{equation} Thus the intertwinig property of $ \BU^{(m_1,m_2,\dots,m_M)}(z)$ is established. \section{Spectrum and eigenspaces of the operators $\Xi^{(n)}(\omega) $ $ (n=1,\dots,N-1)$ forming the hierarchy of $U$-invariant Spin Models.} \subsection{Eigenspaces $H^{\mu}_{\bf{B}}(\omega)$ of the generating function $\Xi(u;\omega) $} {\bf 1.} Let $\mu \equiv (m_1,m_2,\dots,m_M) \in {\frak M}_N .$ Define a subspace $H^{\mu}_{{\cal B}}(\omega) \subset H_{{\cal B}}(\omega) \subset H $ by applying the evaluation map $Ev(\omega)$ (Cf. {\bf 3.1}) to the ``bosonic'' subspace $ {\cal B}^{(m_1,m_2,\dots,m_M)} $ introduced in the previous section: \begin{equation} H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) := Ev(\omega){\cal B}^{(m_1,m_2,\dots,m_M)}\quad ((m_1,m_2,\dots,m_M) \in{\frak M}_N). \end{equation} {}From (3.3.13,.14,.19) we obtain \begin{multline} \Xi(u;\omega) H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) = \left(\sum_{i=1}^M (\delta^{(m_i)}_1(u) - \theta(u)m_i)\right)H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) \\ ((m_1,m_2,\dots,m_M) \in{\frak M}_N). \end{multline} Where $ \delta^{(m_i)}_1(u) $ was defined in (4.4.6) and $\theta(u)$ was defined in (2.2.6). The last equation says that $ H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) $ is an eigenspace of $\Xi(u;\omega)$ unless $ H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) \equiv 0 $. \mbox{} \noindent {\bf 2.} Application of the evaluation map $ Ev(\omega) $ to ${\cal B}^{(m_1,m_2,\dots,m_M)}$ (5.5.19) yields the explicit expression for the space $H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega)$: \begin{equation} H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) = \BU^{(m_1,m_2,\dots,m_M)}(\omega){\cal Z}^{(m_1,m_2,\dots,m_M)}\quad ((m_1,m_2,\dots,m_M) \in{\frak M}_N). \end{equation} Where \begin{multline} \BU^{(m_1,m_2,\dots,m_M)}(\omega) := \BU^{(m_1,m_2,\dots,m_M)}(z_1 =\omega^1,\dots,z_N = \omega^N) = \\ = (-(q^2+1))^M \sum_{\sigma\in S^{(m_1,m_2,\dots,m_M)}_{N,(m_1,m_2,\dots,m_M)}}\varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}(\omega){\Bbb Y}(\sigma) \Pi^{-,(m_1,m_2,\dots,m_M)}_{\{1,2,\dots,M \}} \; , \\ \BU^{(m_1,m_2,\dots,m_M)}(\omega) : H \mapsto H . \end{multline} To get the expression for $\BU^{(m_1,m_2,\dots,m_M)}(\omega)$ we used the Corollary 1 to Proposition 9. \mbox{} \noindent Let us introduce the notation \begin{equation} W^{(m_1,m_2,\dots,m_M)} := \Pi^{-,(m_1,m_2,\dots,m_M)}_{\{1,2,\dots,M \}}{\cal Z}^{(m_1,m_2,\dots,m_M)}\quad ((m_1,m_2,\dots,m_M) \in{\frak M}_N). \end{equation} The definition of ${\cal Z}^{(m_1,m_2,\dots,m_M)}$ given in (5.5.17) leads to a more explicit form of $W^{(m_1,m_2,\dots,m_M)}$: \begin{multline} W^{(m_1,m_2,\dots,m_M)} = \\ S_q(\underset{1}{V}\otimes\dots\otimes\underset{m_1-1}{V})\otimes A_q(\underset{m_1}{V}\otimes\underset{m_1+1}{V})\otimes S_q(\underset{m_1+2}{V}\otimes\dots\otimes\underset{m_2-1}{V})\otimes A_q(\underset{m_2}{V}\otimes\underset{m_2+1}{V})\otimes\ \\ \ldots \otimes S_q(\underset{m_M+2}{V}\otimes\dots\otimes\underset{N}{V}) \; \subset \; H := \underset{1}{V}\otimes\underset{2}{V}\otimes\ldots\otimes\underset{N}{V}. \end{multline} Where $A_q$ signifies $q$-antisymmetrization (1.1.15). Observe that $W^{(m_1,m_2,\dots,m_M)}\quad ((m_1,m_2,\dots,m_M) \in{\frak M}_N)$ is an irreducible highest-wight $U$-module with the $U$-action is given by \begin{equation} L_a(u;\{q^{l_{\sigma[0]_i}}\}):= L_{a1}(uq^{l_{\sigma[0]_1}})L_{a2}(uq^{l_{\sigma[0]_2}})\dots L_{aN}(uq^{l_{\sigma[0]_N}}) \; \in \; End(H). \end{equation} Where we used the notation \begin{equation} \sigma[0] := (m_1,m_1+1)(m_2,m_2+1)\ldots (m_M,m_M+1) \quad ((m_1,m_2,\dots,m_M) \in{\frak M}_N). \end{equation} The Drinfeld polynomial of this module is \begin{equation} Q^{(m_1,m_2,\dots,m_M)}(u) = \prod\begin{Sb}1\leq n \leq N \\ n \neq m_i,m_i+1 \end{Sb}(1 - q^{-l_n}u). \end{equation} \mbox{} \noindent According to (6.6.3) we have \begin{multline} H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) = \Check{\BU}^{(m_1,m_2,\dots,m_M)}(\omega)W^{(m_1,m_2,\dots,m_M)}\quad ((m_1,m_2,\dots,m_M) \in{\frak M}_N), \\ \Check{\BU}^{(m_1,m_2,\dots,m_M)}(\omega):= (-(q^2+1))^M \sum_{\sigma\in S^{(m_1,m_2,\dots,m_M)}_{N,(m_1,m_2,\dots,m_M)}}\varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}(\omega){\Bbb Y}(\sigma). \end{multline} \mbox{} \noindent {\bf 3.} The space $ H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) \quad ((m_1,m_2,\dots,m_M) \in{\frak M}_N)$ is a $U$-module with the $U$-action given by $T_a^0(u;\omega)$ defined in (3.3.19). Explicitely (Cf. 5.5.23): \begin{multline} T_a^0(u;\omega) H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) = T_a^0(u;\omega)\Check{\BU}^{(m_1,m_2,\dots,m_M)}(\omega)W^{(m_1,m_2,\dots,m_M)} = \\ =(-(q^2+1))^M \sum_{\sigma\in S^{(m_1,m_2,\dots,m_M)}_{N,(m_1,m_2,\dots,m_M)}}\varphi_{\sigma}^{(m_1,m_2,\dots,m_M)}(\omega) L_a(u;\{q^{l_{\sigma_i}}\}){\Bbb Y}(\sigma), \\ L_a(u;\{q^{l_{\sigma_i}}\}):= L_{a1}(uq^{l_{\sigma_1}})L_{a2}(uq^{l_{\sigma_2}})\dots L_{aN}(uq^{l_{\sigma_N}}). \end{multline} Applying (5.5.24) we find that $ \Check{\BU}^{(m_1,m_2,\dots,m_M)}$ is an intertwiner of the modules $H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega)$ and $ W^{(m_1,m_2,\dots,m_M)}$: \begin{multline} T_a^0(u;\omega) H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) = T_a^0(u;\omega)\Check{\BU}^{(m_1,m_2,\dots,m_M)}(\omega)W^{(m_1,m_2,\dots,m_M)} = \\ \Check{\BU}^{(m_1,m_2,\dots,m_M)}(\omega)L_a(u;\{q^{l_{\sigma[0]_i}}\})W^{(m_1,m_2,\dots,m_M)} \quad ((m_1,m_2,\dots,m_M) \in{\frak M}_N). \end{multline} Since $ W^{(m_1,m_2,\dots,m_M)}$ is irreducible so is $H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega)$. The highest-weight vector of $ W^{(m_1,m_2,\dots,m_M)}$ is (Cf. 6.6.6) \begin{multline} \tilde{\Omega}^{(m_1,m_2,\dots,m_M)} := \\ v^+\otimes\dots\otimes v^+\otimes(\underset{m_1}{v^+}\otimes\underset{m_1+1}{v^-}-q v^-\otimes v^+)\otimes \quad\ldots \quad\otimes v^+\otimes\dots\otimes v^+. \end{multline} \noindent If the vector $\Omega^{(m_1,m_2,\dots,m_M)} := \Check{\BU}^{(m_1,m_2,\dots,m_M)}\tilde{\Omega}^{(m_1,m_2,\dots,m_M)} \in H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega)$ is not zero, it is the highest-weight vector, and the modules $H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega)$ and $ W^{(m_1,m_2,\dots,m_M)}$ are isomorphic , with $\Check{\BU}^{(m_1,m_2,\dots,m_M)}$ defining the isomorphism explicitely. \mbox{} \noindent {\bf 4.} In order to show that $\Omega^{(m_1,m_2,\dots,m_M)} \quad ((m_1,m_2,\dots,m_M) \in{\frak M}_N) $ is not zero we compute this vector at $q =0$. \noindent Consider the matrix ${\Bbb Y}(\sigma)\quad (\sigma \in S_{N,(m_1,m_2,\dots,m_M)}^{(m_1,m_2,\dots,m_M)})$ that enters the definition of $\Check{\BU}^{(m_1,m_2,\dots,m_M)}$. According to (5.5.15,.16): \begin{align} {\Bbb Y}(\sigma) & = {\mathrm {Id}} \quad ( \sigma = \sigma[0]:= (m_1,m_1+1)\dots(m_M,m_M+1)), \\ {\Bbb Y}(\sigma) & = {\Bbb Y}(\sigma)' Y^-_{i,i+1}(q^{l_{\sigma[0]_i}-l_{\sigma[0]_{i+1}}}) \quad ( \sigma \neq \sigma[0]). \end{align} Where depending on $\sigma$ , $i$ takes one of the values in the set $ \{m_k-1,m_k+1\}_{k\in\{1,2,\dots,M \}}$. For any such $i$ we have $ \sigma[0]_i - \sigma[0]_{i+1} = -2 .$ ${\Bbb Y}(\sigma)'$ is either identity or a product of intertwiners of the form $ Y^-_{j,j+1}(q^{-2r}) \quad j\in (\{1,\dots,N-1\})$ where $ r \geq 2 . $ \noindent For $r \geq 2$ we find \begin{equation} Y^-_{j,j+1}(q^{-2r})|_{q = 0} = -\Pi_{j,j+1}^-(0) = -(|+-><+-|)_{j,j+1} \quad (j=1,\dots,N-1). \end{equation} Thus \begin{multline} \lim_{q\rightarrow 0}{\Bbb Y}(\sigma) = -{\Bbb Y}(\sigma)'|_{q=0}\Pi_{i,i+1}^-(0) \quad (i\in \{m_k-1,m_k+1\}_{k\in\{1,2,\dots,M \}}), \\ \quad (\sigma \in S_{N,(m_1,m_2,\dots,m_M)}^{(m_1,m_2,\dots,m_M)},\; \sigma\neq \sigma[0] ). \end{multline} The highest-weight vector $ \tilde{\Omega}^{(m_1,m_2,\dots,m_M)}$ in the limit $ q\rightarrow 0 $ is \begin{multline} \tilde{\Omega}^{(m_1,m_2,\dots,m_M),q=0} := \\ v^+\otimes\dots\otimes\underset{m_1}{v^+}\otimes\underset{m_1+1}{v^-}\otimes v^+\dots \otimes\underset{m_2}{v^+}\otimes\underset{m_2+1}{v^-} \otimes\quad\ldots \quad\otimes v^+\otimes\dots\otimes v^+. \end{multline} Therefore taking into account (6.6.17) and (4.4.26) we arrive at the following expression for $\Omega^{(m_1,m_2,\dots,m_M)} \in H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega)$ at $q=0$: \begin{multline} \Omega^{(m_1,m_2,\dots,m_M),q=0} = (-1)^M \varphi_{\sigma[0]}^{(m_1,m_2,\dots,m_M)}(\omega) \tilde{\Omega}^{(m_1,m_2,\dots,m_M),q=0} = \\ =\omega^{\frac{1}{2}\sum_{i=1}^M m_i(m_i+1)} \tilde{\Omega}^{(m_1,m_2,\dots,m_M),q=0}. \end{multline} Since $\Omega^{(m_1,m_2,\dots,m_M),q=0}$ is not zero we can argue that same holds for any generic value of $q$ (not a root of unity) and therefore $ H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) $ and $ W^{(m_1,m_2,\dots,m_M)} $ are isomorphic $U$-modules for any generic $q$ and $ (m_1,m_2,\dots,m_M) \\ \in {\frak M}_N .$ \mbox{} \noindent {\bf 5.} From consideration of the case $q=0$ we deduce that \begin{equation} H = \bigoplus_{(m_1,m_2,\dots,m_M) \in {\frak M}_N} W^{(m_1,m_2,\dots,m_M)}. \end{equation} Since the $U$-modules $ H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega)$ have different Drinfeld polynomials (6.6.9) for different motifs $(m_1,m_2,\dots,m_M)$ any two of these modules do not intersect except at zero vector. Therefore we can take the direct sum of all these modules. Since ${\mathrm{ dim}}H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) = {\mathrm {dim}}W^{(m_1,m_2,\dots,m_M)} $, we conclude from (6.6.20) that \begin{equation} H = \bigoplus_{(m_1,m_2,\dots,m_M) \in {\frak M}_N} H^{(m_1,m_2,\dots,m_M)}_{{\cal B}}(\omega) . \end{equation} Thus we have found the complete decomposition of the space of states into eigenspaces of the operator $\Xi(u;\omega)$, as well as the $U$-representation content of this decomposition.
\section{Introduction} Supersymmetry \cite{BaWe} provides a unified framework for the discussion of bosons and fermions and has been proposed as a mechanism to improve the behavior of quantum field theory divergences. It may offer the only viable alternative at present for grand unified models \cite{Ca}. Because it allows to write the Hamiltonian in terms of its ``square roots'' it may provide a new perspective on the constraint equations \cite{DeHaOb,GrCs}. Supergravity was cast in a canonical fashion several years ago by Deser, Kay and Stelle \cite{DeKaSt} , Pilati \cite{Pi} and D'Eath \cite{De}. Teitelboim \cite{TeTa} pointed out that it provided a square root of the Wheeler-DeWitt equation. Jacobson \cite{Ja} showed that new canonical variables similar to the ones introduced by Ashtekar in general relativity could be used in supergravity. Here we will follow this latter approach. In this paper we will show that several ideas that have appeared in the context of the loop quantization of general relativity in terms of the Ashtekar new variables \cite{As} find a natural counterpart in supergravity. There are already partial results concerning the use of new variables and loops in the context of connection representations for supergravity. F\"ul\"op \cite{Fu} first noticed that the theory could be cast in terms of an $GSU(2)$ connection and discussed the Chern-Simons form as a possible state. This state had already been considered in a semi-classical context by Sano and Shiraishi \cite{SaSh}. Matschull \cite{Ma} noted that bosonic Wilson loops were spurious solutions to all the constraints of supergravity. In this paper we will use $GSU(2)$ Wilson loops with nontrivial fermionic content as an overcomplete basis in terms of which we can expand quantum state. We discuss the exponential of the $GSU(2)$ Chern-Simons form as a diffeomorphism invariant solution to all the constraints of supergravity with a cosmological constant with non-trivial fermionic content. We will also build a loop representation for the theory and discuss the kinematical state space. We will see that the loop counterpart of the Chern-Simons solution is associated with the Dubrovnik version of the Kauffman polynomial, which is therefore compatible with the $GSU(2)$ Mandelstam identities. \section{Quantum supergravity in terms of a $GSU(2)$ connection} When cast in a canonical form in terms of the Ashtekar new variables \cite{As}, supergravity is described by variables $\tilde{E}^a_i,A_a^i,\psi_a^A,\tilde{\pi}^a_A$ where lowercase indices from the beginning of the alphabet are spatial tensor indices, lowercase indices from the middle of the alphabet are $SU(2)$ indices and uppercase indices are spinor indices. The $E$'s are densitized triads, the $A$'s are the Sen connection, the $\psi$'s are Grassman-valued Rarita-Schwinger fields and the $\pi$'s are their canonically conjugate momenta. We refer the reader to the paper by Jacobson \cite{Ja} for details. We will follow the notation of F\"ul\"op \cite{Fu}. The canonical framework (with a cosmological constant) has the following constraints, \begin{eqnarray} {\cal G}_i &=& D_a \tilde{E}^a_i + {i\over \sqrt{2}} \pi^a_A \psi_{aB} \sigma^{AB}_i=0 \label{gauss}\\ {\cal S}_A &=& D_a \pi^a_A -i \alpha \tilde{E}^a_i \sigma_{iA}^B \psi_{aB}\\ \bar{\cal S}^A &=& \epsilon^{ijk} \tilde{E}^a_i \tilde{E}^b_j {\sigma_k}^A_B (-4 i D_{[a} \psi_{b]}^B+\sqrt{2}\bar{\alpha} \ut{\epsilon}_{abc} \pi^{cB}) =0 \end{eqnarray} where $\sigma^{AB}_i$ are usual Pauli matrices and $\alpha$ and $\bar{\alpha}$ are such that the cosmological constant is given by $\Lambda = -\alpha \bar{\alpha}$. The usual Hamiltonian and diffeomorphism constraints of general relativity (with the corresponding fermionic extra terms) are obtained by taking Poisson brackets of the ${\cal S}$ and $\bar{\cal S}$ constraints. One could construct a quantum representation by considering wavefunctions $\Psi(A,\psi)$ and promoting the above constraints to wave equations. Matschull observed that if one does so and considers purely bosonic wavefunctions consisting of the Wilson loops built with the bosonic connection $A$ along smooth non-intersecting loops, all the contraints are solved. This presented a puzzle, since such wavefunctions are evidently not diffeomorphism invariant. How could they therefore provide a solution to the diffeomorphism constraint? The answer is given by the fact that the commutator of ${\cal S}$ with $\bar{\cal S}$ give expressions for the usual Hamiltonian and diffeomorphism constraint multiplied by the determinant of the spatial metric. Since Wilson loops based on smooth loops are annihilated by the determinant of the metric, they automatically solved the resulting constraints. Here we will proceed in a different way. F\"ul\"op noticed that if one considers the Gauss law and the right supersymmetry generator, they form under Poisson brackets a graded Lie algebra associated with the $GSU(2)$ group \cite{PaRi}, \begin{eqnarray}\label{4} \{{\cal G}_i,{\cal G}_j\} &=& \sqrt{2} \epsilon_{ijk} {\cal G}_k\\ \{{\cal G}_j, {\cal S}_A\} &=&- {i\over \sqrt{2}} (\sigma_j)^B{}_A {\cal S}_B\\\label{6} \{{\cal S}_A,{\cal S}_B\} &=& -i \alpha (\sigma^j)_{AB} {\cal G}_j. \end{eqnarray} In view of this, one can define new $GSU(2)$ variables $\tilde{\bf E}^a_I$, ${\bf A}_a^J$, ${\bf F}_{ab}^K$, \begin{eqnarray} \tilde{\bf E}^a_I &=& ( \tilde{E}^a_i,\tilde{\pi}^a_A)\\ {\bf A}_a^I &=& (A_a^i,\psi_a^A)\\ {\bf F}_{ab}^I &=& (F_{ab}^i,2 D_{[a} \psi_{b]}^A) \end{eqnarray} The uppercase indices from the middle of the alphabet range from 1 to 5 and refer to a basis for the fundamental representation of $GSU(2)$, given by matrices ${\bf G}_I$, \begin{equation} {\bf G}_1 = -{i\over \sqrt{2}} \left(\begin{array}{ccc} 0&1&0\\1&0&0\\0&0&0\end{array}\right),\quad {\bf G}_2 = -{i\over \sqrt{2}} \left(\begin{array}{ccc} 0&-i&0\\i&0&0\\0&0&0\end{array}\right),\quad {\bf G}_3 = -{i\over \sqrt{2}} \left(\begin{array}{ccc} 1&0&0\\0&-1&0\\0&0&0\end{array}\right), \end{equation} \begin{equation} {\bf G}_4 = \sqrt{\alpha\over \sqrt{2}} \left(\begin{array}{ccc} 0&0&1 \\0&0&0\\0&1&0\end{array}\right),\quad {\bf G}_5 = \sqrt{\alpha\over \sqrt{2}} \left(\begin{array}{ccc} 0&0&0\\0&0&1\\-1 &0&0\end{array}\right),\quad {\bf e} = \left(\begin{array}{ccc} 1&0&0\\0&1&0\\0&0&0\end{array}\right), \end{equation} where we have included the matrix ${\bf e}$ such that $({\bf G}_i)^2=-{\bf e}/2$ for future purposes. The matrices satisfy the commutation relations of the algebra (\ref{4}-\ref{6}). The $GSU(2)$ group has a Killing-Cartan metric associated with the orthosymplectic form \begin{equation} g_{IJ} \lambda^I \lambda^J = \delta_{ij} \lambda^i \lambda^j -\sqrt{2} \alpha \eta_{AB} \lambda^A \lambda^B \end{equation} where the $\lambda$'s are parameters in the group with bosonic components $\lambda_i$ and anticommuting components $\lambda_{A}$ and $\eta_{AB}$ is the antisymmetric symbol in two dimensions. One has the usual normalization of the generators of $GSU(2)$ in the case $\alpha =\sqrt{2}$, the other cases correspond to a constant rescaling of the generators. A remarkable fact is that in terms of the $GSU(2)$ connection one can introduce a covariant derivative ${\bf D}_a$ such that the Gauss law and the right supersymmetry constraint can be written as a single $GSU(2)$ Gauss law \cite{Fu}, \begin{equation} {\bf D}_a \tilde{\bf E}^a_I =0. \end{equation} The left supersymmetry constraint can be written in terms of these variables but it is not invariant under $GSU(2)$ transformations. This is reasonable since it has a nontrivial Poisson bracket with the right generator, giving rise to the usual Hamiltonian and diffeomorphism constraints. In terms of these variables the left supesymmetry constraint can be written as, \begin{equation} \bar{\cal S}_A = {\bf g}_A{}^{IJ}{}_K (\tilde{\bf E}^a_I \tilde{\bf E}^b_J {\bf F}_{ab}^K +i \Lambda \ut{\epsilon}_{abc} \tilde{\bf E}^a_I \tilde{\bf E}^b_J \tilde{\bf E}^{cK}) \end{equation} where \begin{equation} {\bf g}_A{}^{ij}{}_B \equiv -2 i \epsilon^{ij}_k \sigma^k{}_{AB} \end{equation} and all other components vanish. The definition of $\bf g$ is obviously non $GSU(2)$ covariant. The point of writing the equations in terms of these variables is that it allows to find in a straightforward manner several solutions to all the quantum constraint equations. Some of these solutions have a quite nontrivial form when decomposed in terms of the original $SU(2)$ variables. In order to see this let us consider a quantum representation in which wavefunctions are functionals of the $GSU(2)$ connection $\Psi({\bf A})$. The operator $\hat{\tilde{\bf E}}^a_I$ is represented as a functional derivative whereas the operator $\hat{\bf A}_a^I$ is multiplicative. It is immediate to find solutions to the $GSU(2)$ Gauss law. Any $GSU(2)$ invariant expression will do. In particular one can consider Wilson loops constructed with the $GSU(2)$ connection, \begin{equation} W_\gamma ={\rm STr P}\exp\left(\oint_\gamma dy^a {\bf A}_a\right) \end{equation} where $\gamma$ is a loop on the spatial manifold. The trace taken in the above expression is the supertrace \cite{DeWi}, which for any $GSU(2)$ matrix $A$ is given by ${\rm STr}(A)=A_{11}+A_{22}-A_{33}$. Since it is $GSU(2)$ invariant, the Wilson loop is annihilated by the Gauss law. These states are therefore invariant under right supersymmetry transformations and triad rotations. If one studies the action of the left supersymmetry constraint on them it is immediate to notice that if one considers loops $\gamma$ that are smooth (they do not have kinks or intersections), the constraint annihilates these states. The reasons are the same as the ones that made the usual Ashtekar Hamiltonian constraint annihilate a Wilson loop based on a smooth loop \cite{JaSm}. The states have a quite nontrivial fermionic content, as we will discuss in the next section (they correspond to an infinite superposition of terms built with holonomies with arbitrarily high number of fermionic insertions along the loop, if $\alpha\neq0$). Notice that these states are different from those of Matschull \cite{Ma} which were purely bosonic. They reduce to them in the case $\alpha=0$. They also share with those states the pathology that Matschull pointed out: although they solve all the constraints they are not diffeomorphism invariant. We will however find that related states are useful for the construction of a loop representation in the next section. It is possible to find another exact state that solves all the constraint equations. This state is genuinely diffeomorphism invariant and is associated with a non-degenerate spatial metric (with non-vanishing determinant). If one considers the state built by taking the exponential of the Chern-Simons form of the $GSU(2)$ connection, \begin{equation} \Psi_\Lambda({\bf A}) = \exp({i\over 2 \Lambda} \int d^3x {\rm STr}({\bf A}_a \partial_b {\bf A}_c + {\sqrt{2} i \over 3}{\bf A}_a {\bf A}_b {\bf A}_c) \tilde{\epsilon}^{abc} \end{equation} it is immediate to see that it satisfies all the supersymmetry constraints. It is annihilated by the $GSU(2)$ gauss law since it is $GSU(2)$ invariant. It is annihilated by the left supersymmetry constraint for the same reason the corresponding state was annhilated by the Hamiltonian constraint of quantum gravity with a cosmological constant \cite{BrGaPu}: the state has the property $\hat{\tilde{\bf E}}^a_I \Psi(A) ={i \over 2\Lambda} \tilde{\epsilon}^{abc}{\bf F}_{bcI} \Psi(A)$ and therefore the two terms in the left supersymmetry constraint are identical and of opposite sign, cancelling each other. Being the integral of a three form, the state is naturally diffeomorphism invariant. Some concerns may be raised about factor ordering since in the ordering with the triads to the left (where the state is a solution) the constraint that usually corresponds to diffeomorphisms formally does not generate that symmetry. As was discussed in \cite{BrGaPu} using a symmetric regularization for the constraint it actually generates diffeomorphisms and annihilates the state and similar considerations apply here. This state was first introduced by Kodama \cite{Ko} for usual gravity and was discussed in a semiclassical supergravity context by Sano and Shiraishi \cite{SaSh}. F\"ul\"op \cite{Fu} noticed that it was an exact solution of the Wheeler-DeWitt equation of full supergravity. All these discussions were in terms of the $SU(2)$ variables. By writing the state and the constraints in terms of the $GSU(2)$ variables we notice here that it is very easy to see why it is annihilated by the constraints. It will also be the key to finding a counterpart of this state in the loop representation. The Chern-Simons form has been observed to be a solution of all the constraints of Maxwell theory, Yang-Mills \cite{Ja83}, Einstein \cite{BrGaPu}, Einstein-Maxwell \cite{GaPu} and also supergravity. It is remarkable that so different theories have a similar state in common. \section{Loop representation} The usual starting step to construct a loop representation \cite{GaTr,RoSm90} for a theory based on a connection is to expand the states of the theory in terms of a basis of functions given by the Wilson loops constructed with the connection. Such basis is gauge invariant and the resulting representation (the loop representation) is therefore well suited for the description of gauge invariant operators. Crucial to this construction is the knowledge that every gauge invariant quantity can be expanded in terms of Wilson loops. For the bosonic case this last statement is the content of a theorem by Giles \cite{Gi}. For the supersymmetric case we do not know if a similar theorem holds. Even if a theorem like the one above holds, in general one can only expand wavefunctions in terms of products of Wilson loops. In many theories one can re-express an arbitrary product of Wilson loops as a linear combination of products of a fixed minimal number of Wilson loops. The resulting loop representation is based on wavefunctions depending on multiloops of at most that minimal number, through the relation, \begin{equation} \Psi(\gamma_1,\gamma_2,\ldots,\gamma_n) = \int dA W_{\gamma_1}[A] W_{\gamma_2}[A]\cdots W_{\gamma_n}[A] \Psi[A] \end{equation} where $n$ is the minimal number of Wilson loops in terms of which one can express an arbitrary product of Wilson loops. To address the two issues mentioned above, namely if Wilson loops are enough to represent any state and which is the minimal number of Wilson loops needed, let us discuss some properties of the supesymmetric Wilson loops. Because they are traces of group elements, the Wilson loops satisfy certain identities called Mandelstam identities \cite{GaTr86} which reflect particular properties of the group in question. For the group we are considering they are rather nontrivial, so we start with a discussion for the case with $\Lambda=0$. They are given by, \begin{eqnarray} W_{\gamma_1\circ \gamma_2} &=& W_{\gamma_2\circ \gamma_1},\\ W_{\gamma_1} W_{\gamma_2} &=& W_{\gamma_1\circ\gamma_2} + W_{\gamma_1\circ\gamma_2^{-1}} - W_{\gamma_1}-W_{\gamma_2}+1.\label{mandelfund} \end{eqnarray} These identities allow to express any product of Wilson loops as a linear combination of Wilson loops. They can be combined in many nontrivial ways\footnote{For the usual loop representation of quantum gravity, which is based on an $SU(2)$ group, the identity (\ref{mandelfund}) reads $ W_{\gamma_1} W_{\gamma_1} = W_{\gamma_1\circ\gamma_2} + W_{\gamma_1\circ\gamma_2^{-1}}$.}. For instance, it follows from (\ref{mandelfund}) taking $\gamma_1$ to a point that, \begin{equation} W(\gamma) = W(\gamma^{-1}). \end{equation} In order to derive these identities one needs to recall that a generic element of the group in the case $\Lambda=0$ is written as ${\bf 1} +\phi_0 {\bf e} + \phi_I {\bf G}^I$ with $(1+\phi_0)^2+\sum_{I=1}^3 \phi_I^2 =1$. For the case $\Lambda\neq 0$ it is harder to find the identities. We have only suceeded in finding part of them. In order to do this we make use of the techniques introduced in \cite{UrWaZe} that related the Mandelstam identities with the Cayley-Hamilton theorem. This theorem states that any matrix is a root of its characteristic polynomial. For supermatrices a generalization of this statement has been discussed in reference \cite{UrMo}. There it was established that a supermatrix is a root of a polynomial associated with the characteristic polynomial, which acts as a generalization of the Cayley-Hamilton polynomial to the case of supermatrices. For a generic matrix of the orthosymplectic group (which includes as a subgroup $GSU(2)$) the Cayley-Hamilton equation is \cite{UrWaZe}, \begin{equation}\label{21} {\bf H}^3(\gamma) -({\rm STr}({\bf H}(\gamma))+2) ({\bf H}^2(\gamma)-{\bf H}(\gamma))-1 =0, \end{equation} If one now multiplies it by ${\bf H}(\eta) {\bf H}(\gamma)^{-1}$ and takes the trace one gets, \begin{eqnarray} &&{\rm STr}({\bf H}(\eta){\bf H}(\gamma)^2))- ({\rm STr}({\bf H}(\gamma))+2) {\rm STr}({\bf H}(\eta){\bf H}(\gamma))+\nonumber\\ &&\quad({\rm STr}({\bf H}(\gamma))+2) ({\rm STr}({\bf H}(\eta))) -{\rm STr}({\bf H}(\eta){\bf H}(\gamma)^{-1})=0.\label{mandid} \end{eqnarray} This is a ``Mandelstam identity'' in the sense that it is satisfied by the traces of the holonomy we are considering. However one is usually interested in identities that allow to reduce the number of traces. {}From this point of view, the above identity does not help. One can derive another identity that allows to reduce the number of traces in a product by considering the generalization of formula (21) to the case of generic $(2,1)$ matrices\footnote{An arbitrary $3\times 3$ matrix in which the $13$, $23$, $31$ and $32$ components are Grassmanian.} \cite{BeUr}. One then considers a $(2,1)$ matrix given by a sum of $GSU(2)$ matrices and inserts it in the identity. The resulting expression relates products of traces of the $GSU(2)$ matrices. and allows to expand an arbitrary product of Wilson loops as a linear combination of products of three Wilson loops. When we build the loop representation it will therefore be based on wavefunctions depending at most on three loops. There is yet another Mandelstam identity that has to be considered, that which reflects the fact that the matrices are of unit determinant. We will not consider it here. The use of Cayley-Hamilton techniques allows us to state that in general, for any theory based on a group or supersymmetric group the Cayley-Hamilton theorem can be used to prove that the loop representation will involve wavefunctions of a finite number of loops. Let us address the question of up to what extent one can use combinations of these Wilson loops to represent any gauge invariant state. As we mentioned before, in the bosonic case this was ensured by Giles' theorem. Here we do not have a similar theorem. However it is easy to see that it is likely to be a problem. Consider the case $\Lambda=0$. As we saw, the supersymmetric Wilson loops reduce to a purely bosonic expression coinciding with the Wilson loop constructed with the bosonic $SU(2)$ connection. Therefore they clearly fail to capture any fermionic information. For the $\Lambda\neq 0$ case we do not know what the situation exactly is. It should be noted that the $\Lambda=0$ case is somewhat pathological from the point of view of the $GSU(2)$ symmetry we are exploiting since the invariant orthosymplectic form collapses to the usual Euclidean form. We will proceed to build a pure loop representation to highlight other aspects of the construction but it should be forewarned that it is possible that the resulting quantum representation only captures part of the information present in the theory. Apart from the above mentioned peculiarities of the supersymmetric case, there is yet another difference with the usual loop representation of a bosonic theory. The $GSU(2)$ Wilson loops naturally implement the symmetry of the theory under triad rotations and right supersymmetry transformations. If we construct a loop representation using them as a basis for states it will be difficult to write an expression in such a representation for the left supersymmetry generator $\bar{\cal S}$, which is not invariant under $GSU(2)$ rotations. It is like trying to represent in the usual loop representation for a Yang-Mills theory a non-gauge invariant quantity. A similar situation arises in gravity when one considers the diffeomorphism constraint. The space of solutions to this constraint is given by functions of knots. In this space we cannot represent the Hamiltonian constraint, since it is not diffeomorphism invariant. This forces us to define the Hamiltonian constraint on functions of loops rather than on functions of knots, in spite of the fact that the space of solutions to the constraint that are physical will be given by functions of knots. Here we will proceed in a similar way. We will build a representation in terms of loops and an extra structure, in order to be able to represent the left supersymmetry constraint. The solutions to the constraint however, will be given by pure functions of loops. In order to accomplish this let us examine in detail the expression for the $GSU(2)$ Wilson loop we introduced above. If one expands out the expression for the path ordered exponential one has that, \begin{equation} W_\gamma({\bf A}) = 1 +\sum_{n=1}^\infty \oint_\gamma dy_1^{a_1} \cdots \int_o^{y_{n-1}} dy_n^{a_n} {\rm Tr}({\bf A}_{a_1}(y_1)\ldots {\bf A}_{a_n}(y_n)). \end{equation} To further discuss this expression we introduce the following notation: the non-boldface matrix $A$ will represent the combination ${\bf A}_a^i G_i$, $i=1\ldots 3$ and the matrix $\psi_a$ will represent the combination ${\bf A}_a^B G_B$, $i=4,5$. Both are $GSU(2)$ matrices. In terms of these quantities the above expression can be written as, \begin{eqnarray} W_\gamma({\bf A}) &=& 1+\sum_{n=1}^\infty \oint_\gamma dy_1^{a_1} \cdots \int_o^{y_{n-1}} dy^{a_n}_n {\rm STr}(A_{a_1}(y_1)\ldots \ldots A_{a_n}(y_n))\nonumber\\ &&+\sum_{n=2}^\infty \sum_{i=1}^n \oint_\gamma dy_1^{a_1} \cdots \int_o^{y_{n-1}} dy^{a_n}_n{\rm STr}(A_{a_1}(y_1)\ldots A_{a_{i-1}}(y_{i-1}) \psi_{a_i}(y_i) \ldots A_{a_n}(y_n))\nonumber\\ &&+\sum_{n=2}^\infty \sum_{j=2}^n \sum_{i=1}^{j-1} \oint_\gamma dy_1^{a_1} \cdots \int_o^{y_{n-1}} dy^{a_n}_n\\&&\quad\times{\rm STr}(A_{a_1}(y_1)\ldots A_{a_{i-1}}(y_{i-1}) \psi_{a_i}(y_i) \ldots A_{a_{j-1}}(y_{j-1}) \psi_{a_j}(y_j) \ldots A_{a_n}(y_n)). \nonumber \end{eqnarray} The above expression can be rearranged as, \begin{eqnarray} W_\gamma({\bf A}) &=& \sum_{n=0}^\infty GT^n_{\gamma}({\bf A})\label{wila}\\ GT^n_{\gamma}({\bf A})&\equiv& \oint_\gamma dy_1^{a_1} \cdots \int_o^{y_{n-1}} dy^{a_n}_n {\rm Tr}(\psi_{a_1}(y_1) U(\gamma_{y_1}^{y_2}) \psi_{a_2}(y_2) \ldots U(\gamma_{y_{n}}^{y_1})). \end{eqnarray} Each of the terms in the summation in (\ref{wila}) is $SU(2)$ invariant. In order to see this notice that the Gauss law (\ref{gauss}) acts homogeneously in $\psi$ and therefore does not mix the different terms in (\ref{wila}). Since the sum is invariant, each term has to be invariant as well. This can be explicitly checked and one notices that it occurs in a rather nontrivial fashion, using explicitly the fact that the $\psi$'s are Grassman-valued. These invariants can be viewed as holonomies that one breaks at certain points and inserts a $\psi$ and then integrates the point along the loop. These invariants are quite different than the ones one usually considers when building loop representations for theories coupled to fermions, where the fermions can only appear at the ends of open paths and is directly connected with the $GSU(2)$ invariance that is characteristic of this theory. One can build a loop representation based on the invariants introduced above either through a transform or through the introduction of further invariants that involve the momenta ${\bf E}$. We will choose the first approach for brevity. Given a wavefunction in the connection representation $\Psi({\bf A})$ one constructs a wavefunction in the loop representation by, \begin{equation} \Psi(\epsilon,\gamma) = \sum_{i=1}^\infty \int d {\bf A} \epsilon^n GT^n_{\gamma}({\bf A}) \Psi({\bf A}) \end{equation} Wavefunctions in the loop representation are characterized by a loop and an real parameter $\epsilon$. The $GT^n_{\gamma}$'s satisfy a series of identities similar to the Mandelstam identities for the Wilson loops we introduced at the beginning of this section, we will not discuss them here since we will not need them for the issues addressed in this paper. In this representation the left supersymmetry constraint cannot be directly represented since it is not $SU(2)$ invariant. However one can build very easily $SU(2)$ invariant expressions that are equivalent to the left supersymmetry constraint, for instance, by contracting it with the spin $3/2$ field. The fact that the representation is not completely cast in terms of loops appears to detract from the geometric nature of the usual loop representation. However, as we pointed out, one needs the extra parameter to represent non $GSU(2)$ invariant quantities and states. The solutions to the constraint are $GSU(2)$ invariants and therefore are expressible purely in terms of loops. We will analyze in the following section one such solution, the one that is obtained by transforming into the loop representation the Chern-Simons state we discussed in the previous section. \section{The Chern-Simons state and the Dubrovnik version of the Kauffman Polynomial as a solution to the super-Wheeler--DeWitt equation} Let us consider the expression in the loop representation of the solution to all the constraint equations with a cosmological constant given by the exponential of the super Chern-Simons form, \begin{equation} \Psi(\gamma_1,\ldots,\gamma_n) = \int dA e^{{i\over 2 \Lambda} S_{CS}} W_{\gamma_1}[A] W_{\gamma_2}[A]\cdots W_{\gamma_n}[A] \end{equation} where $S_{CS}$ is the super Chern-Simons form we introduced above. Since it is a $GSU(2)$ invariant state we only need to introduce Wilson loops in the transform. Assuming that one is using a diffeomorphism invariant measure in the transform, the resulting state in the loop representation has to be a knot polynomial. We will show that it is the Kauffman Polynomial, Dubrovnik version. The derivation is analogous to the one giving the Kauffman bracket as a state of ordinary bosonic gravity in the loop representation and was discussed in reference \cite{BrGaPu}. We will first study the transform in the case of a one component loop \begin{equation} \Psi(\gamma) = \int dA e^{{i\over 2 \Lambda} S_{CS}} W_\gamma[A].\label{30} \end{equation} In the case of bosonic gravity, the transform of the Chern-Simons state is given by the HOMFLY polynomial. The skein relations satisfied by a regular-isotopic\footnote{As in the case of the bosonic transform of the Chern-Simons state the resulting invariant is a function of framed links. In the bosonic case that is the reason the result is the HOMFLY polynomial rather than the Jones polynomial.} HOMFLY polynomial $F(z,t,\alpha)_\gamma$ are, \begin{eqnarray} F_U&=&1\\ F_{\hat{L}_+} &=& \alpha F_{\hat{L}_0}\\ F_{\hat{L}_-} &=& \alpha^{-1} F_{\hat{L}_0}\\ t F_{L_+} -t^{-1} F_{L_-} &=& z F_{L_0} \label{skein3} \end{eqnarray} where these relations are to be understood as follows. Given a knot, pick a crossing in its planar diagram and replace it with either $L_+$, $L_-$ or $L_0$ as depicted in figure \ref{skein} or figure \ref{hats} for the hatted elements. Evaluate the polynomial on the resulting links. The resulting polynomials are related by the above expressions. The first equation is a normalization condition stating that the polynomial evaluated on the unknot is equal to one. \begin{figure} \hspace{3.7cm}\epsfxsize=300pt \epsfbox{fig1.eps} \caption{The different crossings involved in the skein relations} \label{skein} \end{figure} \begin{figure} \hspace{2.7cm}\epsfxsize=400pt \epsfbox{fig4.eps} \caption{Regular isotopy invariants are not invariant under the addition and removal of a curl. This is determined by the skein relations involving the elements shown} \label{hats} \end{figure} We will show that the expression of the transform of the super Chern-Simons state satisfies related but different skein relations. In order to do this we will perform the following computation first suggested by Smolin \cite{Sm} and Cotta-Ramusino et al \cite{CoGuMaMi}. Starting from the expression of the transform evaluated at an intersection we will append to it an infinitesimal loop in such a way as to turn the intersection into an over crossing. We will then repeat the same procedure to turn it into an under crossing. We see that the difference of the resulting expressions is related to the value of the transform evaluated at $L_0$ through a skein relation of the same nature of (\ref{skein3}), but for a different polynomial. The variation of the loop transform of the state when a small loop of element of area $\sigma^{ab}$ is appended to a single-component loop $\gamma$ at an intersection as shown in figure \ref{inter} is given by, \begin{figure} \hspace{5.7cm}\epsfxsize=100pt \epsfbox{fig2.eps} \caption{The addition of a small loop at an intersection} \label{inter} \end{figure} \begin{equation} \sigma^{ab} \Delta_{ab}(y) \Psi[\gamma] = -2 \Lambda i (-1)^P \int dA \sigma^{ab} \epsilon_{dab} {\rm STr}[{\bf G}_I {\bf H}_{23}(\gamma_y^y) {\bf H}_{41}(\gamma_y^y)] {\delta \over \delta A_d^I(y)} g^{IJ} {\rm exp}({\textstyle {i\over 2\Lambda}} S_{CS}) \end{equation} where $\Delta_{ab}$ is the loop derivative \cite{GaPubook}, $G_I$ is one of the generators of $GSU(2)$ and we have used $\Delta_{ab}(x) {\rm STr}[{\bf H}(\gamma)] = {\bf F}_{ab}^I(x) {\rm STr}[{\bf G}_I {\bf H}(\gamma_x^x)]$ and $\gamma_x^x$ is the loop with origin at the point $x$. The labels on the holonomy indicate the connectivity at the intersection, for instance $H_{23}$ is the loop that begins at $2$ and ends at $3$ as indicated in the figure. We have also used the fundamental property of the Chern-Simons state that we discussed in section 3, \begin{equation} {\delta \over \delta {\bf A}_a^I} \Psi_{\Lambda}[A] = {i \over 2 \Lambda} \tilde{\epsilon}^{abc} {{\bf F}_{bc}}_I \Psi_{\Lambda}[A] \end{equation} to convert the $F_{ab}$ factor due to the loop derivative into a functional derivative acting on the exponential of the Chern-Simons action. It should be recalled that indices like $I$ are raised and lowered with the orthosymplectic metric $g^{IJ}$. The factor $(-1)^P$ is introduced to take care of the flips in sign that take place when $F_{ab}$ is moved from the left of the supertrace to the right of it before converting it into a functional derivative and is determined by the odd/even nature of each of the components of $F_{ab}$ and the supertrace. Integrating by parts and choosing the element of area $\sigma^{ab}$ parallel to segment 1-2 so that the contribution of the functional derivative corresponding to the action on the segment 1-2 vanishes (since the volume element is zero) we get, \begin{eqnarray} \sigma^{ab} \Delta_{ab} \Psi[\gamma] &=& 2 i \Lambda \int dA \sigma^{ab} \epsilon_{abc} \int dv^c \delta(y-v) \times \nonumber\\ && {\rm STr}[ {\bf G}_I {\bf H}_{23}(\gamma_y^y) {\bf G}_J {\bf H}_{41}(\gamma_y^y) ] g^{IJ} {\rm exp} ({\textstyle {i \over 2 \Lambda}} S_{CS})(-1)^{JH}, \end{eqnarray} and in the integration by parts the $(-1)^P$ factor is cancelled but a factor $(-1)^{JH}$ is introduced when the functional derivative is ``introduced'' in the supertrace at the end of the holonomy going from $2$ to $3$ and is determined by the odd/even nature of the connection in the functional derivative and the components of the holonomy going from $2$ to $3$. As is usual in these kinds of variational derivations \cite{Br}, a regularization of the volume element determined by the element of area of the loop derivative and the tangent to the loop is needed, we take it in such a way that the volume is normalized to be $\pm1$ depending on the orientation. We now make use of the Fierz identity for $GSU(2)$, \begin{eqnarray} g^{IJ} {({\bf G}_I)^{\kappa}}_\beta {({\bf G}_J)^{\gamma}}_\delta &=& {{\bf e}^\kappa}_\delta {{\bf e}^\gamma}_\beta-{1\over 2} {{\bf e}^\kappa}_\beta {{\bf e}^\gamma}_\delta \nonumber\\ &&+{\textstyle {1\over 2}}(\delta^\kappa_1 \delta^3_\beta +\delta^\kappa_3 \delta^2_\beta)(\delta^\gamma_2 \delta^3_\delta-\delta^\gamma_3 \delta^1_\delta) \nonumber\\ &&-{\textstyle {1\over 2}} (\delta^\kappa_2 \delta^3_\beta -\delta^\kappa_3 \delta^1_\beta)(\delta^\gamma_1 \delta^3_\delta+\delta^\gamma_3 \delta^2_\delta) \end{eqnarray} and taking into account the explicit general form of an element of $GSU(2)$ \cite{To}, \begin{equation} \left(\begin{array}{ccc}a (1-\alpha p q/2)&b (1-\alpha p q/2)&\sqrt{\alpha} p\\c (1-\alpha p q/2)&d (1-\alpha p q/2)&\sqrt{\alpha} q\\\sqrt{\alpha} (-a q +c p)& \sqrt{\alpha} (d p -b q)& (1+\alpha p q)\end{array}\right)\label{explicit} \end{equation} where $ad-bc=1$ and $p$ and $q$ are Grassmanian variables one finally gets, \begin{eqnarray} &&\sigma^{ab} \Delta_{ab} \Psi[\gamma] = \label{potskein}\\ &=& i \Lambda \int dA \; \sigma^{ab} \epsilon_{abc} \int dv^c \delta(y-v) {\rm STr}[{\bf H}_{23}(\gamma_y^y)] {\rm STr}[{\bf H}_{41}(\gamma_y^y)] {\rm exp}({\textstyle {i \over 2\Lambda}} S_{CS}) \nonumber\\ &+& i \Lambda \int dA \; \sigma^{ab} \epsilon_{abc} \int dv^c \delta(y-v) {\rm STr}[{\bf H}_{23}(\gamma_y^y) {\bf H}^{-1}_{41}(\gamma_y^y)] {\rm exp}( {\textstyle {i \over 2\Lambda}} S_{CS}).\nonumber \end{eqnarray} At this point we may proceed as in the bosonic case and reinterpret relation ( \ref{potskein}) as a skein relation. The form of the expression we got suggests that the skein relation is, \begin{equation} \Psi[L_\pm] -\Psi[L_I] =\pm i \Lambda ( \Psi[L_0] + \Psi[L_W] ).\label{skeinbuena} \end{equation} There is a subtle difference, however, between expression (\ref{potskein}) and the skein relation (\ref{skeinbuena}). The first involves a rerouting of a portion of the loop. The implication of that rerouting for the connectivity of the loop at the intersection (the only ingredient that participates in the skein relation) is only determined given an initial connectivity of the loop at the intersection. The above skein relation corresponds to starting with a loop with a connectivity at the intersection corresponding to a single loop. One could keep the same intersection and reconnect the strands in such a way that the initial loop has two components. That requires a separate derivation for the addition of an infinitesimal element of area. Let us therefore perform such a calculation. We consider a two component loop with an intersection as shown in figure \ref{multi}, \begin{equation} \Psi(\gamma_1,\gamma_2) = \int dA e^{{i\over 2 \Lambda} S_{CS}} W_{\gamma_1}[A] W_{\gamma_2}[A].\label{42} \end{equation} The addition of a small loop at the intersection gives, through a calculation very similar to the one performed in the case of a single component loop, \begin{eqnarray} \sigma^{ab} \Delta_{ab} \Psi(\gamma_1, \gamma_2) &=& 2 i \Lambda \int dA \sigma^{ab} \epsilon_{abc} \int dv^c \delta(y-v) \times \nonumber\\ && {\rm STr}[ {\bf G}_I {\bf H}_{12}({\gamma_1}_y^y)] \times {\rm STr}[ {\bf G}_J {\bf H}_{43}({\gamma_2}_y^y) ] g^{IJ} {\rm exp} ({\textstyle {i \over 2 \Lambda}} S_{CS})(-1)^{JH_{12}}, \end{eqnarray} and the labels refer to the figure \ref{multi}. \begin{figure} \hspace{3.7cm}\epsfxsize=300pt \epsfbox{fig5.eps} \caption{The same ``straight-through'' crossing can correspond to a two loop link or a single loop depending on the connectivity of the diagram.} \label{multi} \end{figure} Using the Fierz identity this expresion can be written as \begin{eqnarray} &&\sigma^{ab} \Delta_{ab} \Psi(\gamma_1,\gamma_2) = \label{44}\\ &=& i \Lambda \int dA \; \sigma^{ab} \epsilon_{abc} \int dv^c \delta(y-v) {\rm STr}[{\bf H}_{12}({\gamma_1}_y^y) {\bf H}_{43}({\gamma_2}_y^y)] {\rm exp}({\textstyle {i \over 2\Lambda}} S_{CS}) \nonumber\\ &-& i \Lambda \int dA \; \sigma^{ab} \epsilon_{abc} \int dv^c \delta(y-v) {\rm STr}[{\bf H}_{12}({\gamma_1}_y^y) {\bf H}^{-1}_{43}({\gamma_2}_y^y)] {\rm exp}( {\textstyle {i \over 2\Lambda}} S_{CS}).\nonumber \end{eqnarray} These results can be interpreted as the following skein relation for the intersection, \begin{equation} \Psi[L_\pm] -\Psi[L_I] =\pm i \Lambda ( \Psi[L_0]- \Psi[L_W]) \end{equation} for the case of a two component link. If we now replace $W(\gamma)$ by $\hat{W(\gamma)}= -W(\gamma)$ in the above expression (\ref{30},\ref{potskein}) and in (\ref{42},\ref{44}), signs change in such a way that both expressions can be associated with a single skein relation, \begin{equation} \Psi[L_{+}] -\Psi[L_{-}] = -2 i \Lambda ( \Psi[L_0] - \Psi[L_W] ). \end{equation} Some comments are in order. The study of the variational relation obtained by adding an infinitesimal element of area to a crossing with kinks proceeds along similar lines as the one we exhibited above. The main difference is that the volume term has two contributions that add up with opposite signs, corresponding to the addition of a small area along the two different tangents that enter the kink. With a suitable regularization, the whole contribution can be taken to be zero. This implies that an intersection with a kink can be taken as equivalent to no intersection at all, as shown in the figure \ref{kinks}, and this is what allow us to replace the kinks by the $L_0$ and $L_W$ in equations (\ref{potskein},\ref{44}) to obtain the skein relations. \begin{figure} \hspace{3.7cm}\epsfxsize=300pt \epsfbox{fig3.eps} \caption{Intersections without (a) and with (b) kinks. Under the present regularization intersections of type (b) are equivalent to no intersection (c).} \label{kinks} \end{figure} To completely characterize the polynomial, we need the value of the polynomial for the unknot, which is -1, to be consistent with the redefinition of the holonomy as minus the supertrace we introduced above. Also, since the polynomial turns out to be a regular isotopy invariant, we need to study the effect of the addition of a ``curl'' at a point with no intersection. The details are exactly the same as those in reference \cite{BrGaPu} so we will omit them here. The only new ingredient needed is a contraction of the Fierz identity, \begin{equation} g^{IJ} {({\bf G}_I)^\kappa}_\beta {({\bf G}_J)^\beta}_\gamma= \delta^\kappa_\gamma, \end{equation} and one gets as result, \begin{equation} \Psi[\hat{L}_\pm] = (1 \pm 2 \Lambda i) \Psi[\hat{L}_0] \end{equation} where the meaning of the hatted elements is described in figure \ref{hats}. Let us considered now the Dubrovnik version of the Kauffman Polynomial; it is defined by the following relations \cite{onknots}, \begin{eqnarray} \Psi[L_{+}] -\Psi[L_{-}] &=& z ( \Psi[L_0]- \Psi[L_W]) \\ \Psi[{\rm unknot}] &=& (a-a^{-1})/z + 1 \\ \Psi[\hat{L}_{+}] &=& a \Psi[L_0] \\ \Psi[\hat{L}_{-}] &=& a^{-1}\Psi[L_0] \newline \end{eqnarray} One can check that this correponds exactly to the results we obtained to first order provided $z= -2i\Lambda$ and $a = 1+2i\Lambda $ and $ \Psi[L_0] = -1 $ which is exactly the case. We therefore see that the Dubrovnik version of the Kauffman Polynomial bears the same relationship with a $GSU(2)$ Chern--Simons theory as the Kauffman bracket has with the usual $SU(2)$ theory. It is also important to remember that we have shown that the Dubrovnik version of the Kauffman polynomial is the transform into the loop representation of an exact quantum state of supergravity. It therefore should follow (as it happens in ordinary gravity) that the polynomial should be annihilated by all the constraints of quantum supergravity. It is also worthwhile pointing out that one could explore states without cosmological constant based on the ambient invariant polynomial associated with the one we found, as happens in the bosonic case \cite{essay}. An interesting aspect is that Saleur and Zhang \cite{Sa,Zh} have considered representations of the braid group associated with graded groups and found associated knot polynomials. It would be interesting to check if these have a relation with the Dubrovnik version of the Kauffman polynomial, as the calculation we have performed suggests. It would also be important to find ways to compute the skein relations we found to first order in exact form. This could be accomplished via a supersymmetric version of the Moore-Seiberg-Witten \cite{MoSe,Wi} construction based on conformal field theory. Finally, one could evisage computing explicit expressions for each coefficient of the Dubrovnik version of the Kauffman polynomial by perturbatively evaluating the expectation value of a Wilson loop in Chern--Simons theory. For the bosonic case this was first studied by Guadagnini, Martellini, and Mintchev \cite{GuMaMi} and although more complex, the supersymmetric version of this calculation is completely feasible. This will be the first time that explicit expressions for this polynomial have been found. \section{Discussion} This paper explored several issues that arise when trying to construct a loop representation for supergravity using the fact that the theory can be cast as a gauge theory of the $GSU(2)$ group. There are many detailed results that are yet to be derived, as the explicit form of the left supersymmetry constraint in the representation constructed, a complete set of Mandelstam identities and a suitable regularization for the constraint. Yet, we are already able to see the emergence of a rich mathematical structure of the representation to be constructed, in particular concerning the space of physical states of the theory. It is remarkable that a gauge theory with fermions yields a state space that only includes closed loops, contrary to what happens in other cases \cite{GaFo,MoRo}. This may be related to the extra symmetry present in supersymmetric theories equating bosons to fermions. As in the non supersymmetric case it is expected that one could find a basis of gauge invariant states that are free of Mandelstam constraints through the use of spin networks. In this case they would be spin networks associated with a graded group. The properties of such objects are yet to be explored. One could also complete the quantization of the theory in the Euclidean sector using rigorous measure theory, as has been done in the non supersymmetric case. Finally, as a by product we have showed that the knot polynomial associated with Chern-Simons theory based on a graded group is the Dubrovnik version of the Kauffman Polinomial. This is a remarkable result; it allows new insights into that polynomial and opens new perspectives in the search for the conjectured Link Polinomial \cite{onknots2} which has the HOMFLY and the Kauffman Polinomials as particular cases. \acknowledgements We wish to thank Luis Urrutia and Leonardo Setaro for discussions and John Baez for pointing out several references. This work was supported in part by grants NSF-PHY-9423950, NSF-PHY-9396246, NSF-INT-9406269, research funds of the Pennsylvania State University, the Eberly Family research fund at PSU and PSU's Office for Minority Faculty development. JP acknowledges support of the Alfred P. Sloan foundation through a fellowship. We acknowledge support of Conicyt (Uruguay) and Conacyt (Mexico), through grant 4862-E9406.
\section{Introduction} The connection-triad variables introduced by Ashtekar \cite{Ash87} have simplified the constraint equations of quantum gravity; further, these variables suggest that in the future we may be able to reformulate gravity in terms of non-local holonomies rather than local field operators \cite{RovSmo, GambTri}. However, the new variables are unfamiliar, and it is not always clear what they mean physically and geometrically. For example, it is not clear what operators or structures correspond to gravity waves. On the other hand, there has been great progress in constructing diffeomorphism- invariant volume and area operators; these operators turn out to be essentially counting operators for numbers of loops and loop intersections \cite{AV}. The quantum constraint equations are much simpler in the new variables, and solutions to these equations have been found \cite{RovSmo, knotsol} . However, these solutions correspond to some metric, and it is not always clear what that metric is (the problem of physical interpretation again). In order to investigate the metric, one needs a measure, so as to be able to form dot products and take expectation values. Some progress has been made in constructing a measure \cite{measure}. The new approach is not yet truly non-local, and it shares the same renormalization and regularization difficulties which plague other local field theories such as QCD. For gravity, there are some new twists to the old story. The theory is hard to regulate, because the regulators do not always respect diffeomorphism invariance \cite{Blen, OptrOrd, Bor}. However, the theory is astoundingly easy to renormalize \cite{RovSmo}. (Compare to QCD, which is easy to regulate, whereas renormalization is difficult.) Also, it is not always possible to operator-order the gravitational constraints so that both the constraint algebra closes (commutator of two constraints = sum of constraints) and the vector constraints generate spatial diffeomorphisms \cite{knotsol}. (This is a difficulty which the Ashtekar approach shares with more traditional approaches.) The constraint closure is not only essential to the consistency of the Dirac quantization procedure; closure is important even classically. When the 3 + 1 splitup is integrated forward in time, to construct the entire spacetime, the theory will not be invariant under the full four-dimensional diffeomorphism group unless the constraint algebra closes \cite{HKTei}. Thus an invariance of the classical theory is lost if commutator closure is neglected. In this paper we consider the application of the Ashtekar formalism to the problem of plane gravitational waves. "Plane" means the metric possesses two commuting spacelike Killing vectors, and we shall choose coordinates so that these vectors are unit vectors pointing in the x and y directions. \begin{equation} k^{(x)} = \partial _x; k^{(y)} = \partial _y. \label{1.0} \end{equation} We begin the quantization procedure in section 2 by choosing a factor ordering and verifying closure of the constraint algebra. We find a rather surprising result: in the plane wave case, one and the same factor ordering makes the vector constraint into a diffeomorphism generator and allows the algebra to close. This result requires the high symmetry (the two spatial Killing vectors). In section 3 we tackle the problem of constructing a measure. Here we attain some partial success. Our measure respects the reality constraints obeyed by the Ashtekar connections. However, our measure is not invariant under the constraints. The literature on plane waves is vast, but we single out two papers which are especially close to the present paper. Husain and Smolin \cite{HSm} were the first to apply the Ashtekar formalism to the two-Killing vector case. Neville \cite{Nev}, working with the traditional geometrodynamic variables, found the transformations which reduce the Hamiltonian to parameterized free field form and constructed the classical constants of the motion. This paper will be referred to as I. Reference I studied waves which were unidirectional as well as plane, where "unidirectional" in the present coordinates means the waves are moving only in the +z direction. Since unidirectional waves are known to obey a superposition principle, they do not scatter, except off waves moving in the -z direction. The present paper does not use the unidirectional assumption. All results on closure and the measure hold in the presence of scattering. Our notation is typical of papers based upon the Hamiltonian approach with concomitant 3 + 1 splitup. Upper case indices A, B, $\ldots $,I, J, K, $\ldots$ denote local Lorentz indices ("internal" SU(2) indices) ranging over X, Y, Z only. Lower case indices a, b, $\ldots $, i, j, $\ldots $ are also three- dimensional and denote global coordinates on the three-manifold. Occasionally the formula will contain a field with a superscript (4), in which case the local Lorentz indices range over X, Y, Z, T and the global indices are similarly four-dimensional; or a (2), in which case the local indices range over X, Y (and global indices over x, y) only. The (2) and (4) are also used in conjunction with determinants; e.\ g., g is the usual 3x3 spatial determinant, while\ $^{(2)}e$ denotes the determinant of the 2x2 X, Y subblock of the triad matrix $e^A_a$. We use Levi- Civita symbols of various dimensions: $\epsilon _{TXYZ} = \epsilon _{XYZ} = \epsilon _{XY} = +1$. The basic variables of the Ashtekar approach are an inverse densitized triad \E{a}{A} and a complex SU(2) connection \A{A}{a}. \begin {eqnarray} \E{a}{A}& =& \sqrt g e^a_A; \\ \label{eq:1.1} [\E{a}{A},\A{B}{b}]&=& -\hbar \delta (x-x') \delta ^B_A \delta ^a_b. \label{eq:1.2} \end{eqnarray} The local Lorentz indices are vector rather than spinor; strictly speaking the internal symmetry is O(3) rather than SU(2), gauge- fixed to O(2) rather than U(1). \section{Closure of Constraints} Since the result that the constraints do close is surprising, and since any proof of closure is bound to be detailed, it would be helpful if we could state some simple reason why the constraints close, before becoming enmeshed in the details. This is not too hard to do. We consider both the geometrodynamical and Ashtekar approaches, since closure should be independent of choice of basis variables. We begin by establishing some notation. In the coordinate system (1.1), metric components are independent of x and y, and it is possible to bring the metric to a block diagonal form \cite{EhlK}. One 2x2 subblock connects only t and z components of the metric; another 2x2 subbolck connects only x and y components. The x,y subblock may be parameterized using variables suggested by Szekeres \cite{Sz}: \[ ds^2 = e^A [dx^2 e^B \cosh{W} + dy^2 e^{-B} \cosh{W} - 2dxdy \sinh{W} ] \] \begin {equation} + e^{D-A/2} \{ [ -{(\mbox{${\rm N}'$})}^{2} + {(\rm N ^z)}^{2}]dt^2 +2 \rm N^z dzdt + dz^2 \}. \label{eq:2.1} \end{equation} \mbox{${\rm N}'$}\ is not quite the usual lapse N. \begin{equation} \mbox{${\rm N}'$} = \rm N /\sqrt{g_{zz}} . \label{eq:2.2} \end{equation} Now recall the usual reason why the constraints do not close. Let the Hamiltonian density be $\rm N H_0 + \rm N ^i H_i$, with $H_0$ the scalar constraint and $H_i$ the vector constraints. Then the commutator of two scalar constraints, \begin{equation} [\rm N H_0,\rm M H_0] = \delta(x - x')[\rm M\partial_i\rm N - \rm N \partial_i\rm M ]H^j g^{ij}, \label{eq:2.3} \end{equation} does not close because the final inverse metric factor prevents $H_j$ from annihilating the wave funtional, implicitly assumed to stand to the right on both sides of equation~(\ref{eq:2.3}). In the present case, from equation~(\ref{eq:2.1}), the inverse spatial metric is block diagonal, with a 1x1 subblock containing only $g^{zz}$. Also, x and y vector constraints have been gauged away in the process of bringing the metric to block diagonal form, so that the problem contains only a z vector constraint, and the $g^{ij}$ in equation ~\ref{eq:2.3} collapses to $g^{zz} = 1/g_{zz}$. Thus we can completely eliminate the $g^{ij}$ problem by absorbing a factor of $\sqrt{1/g_{zz}} $ into each of the lapse functions N and M, as at equation~(\ref{eq:2.2}). This is the simple reason why the constraints close. The innocuous-looking renormalization~(\ref{eq:2.2}) is the key step. What would the renormalization~(\ref{eq:2.2}) look like in the Ashtekar notation? In that formalism the 3-metric and its conjugate momenta are replaced by complex connection fields \A{A}{a} and densitized inverse triads \E{a}{A}, where A = X,Y,Z is a local Lorentz index. Besides the vector and scalar constraints, there are three new constraints, which generate local SU(2) rotations (or in our case, local O(3) rotations, since A is a vector rather than a spinor index.), The 3x3 \E{a}{A} matrix may be brought to block diagonal form, with one 1x1 subblock plus one 2x2 subblock, exactly as for the 3x3 spatial subblock of the metric $g^{ij}$ The \E{a}{A} matrix is not symmetric, however, so contains more independent components than the metric. To bring the inverse triad matrix to block diagonal form , one must gauge away $9 - (1 + 4) = 4$ inverse triad components, whereas in the metric case one removes only $6 - (1 + 3) = 2$ components. To remove the extra two components one must fix (the x and y vector constraints, as before, plus) two more constraints, the X and Y SU(2) or O(3) generators. The 1x1 subblock of the \E{a}{A} matrix is occupied by \E{z}{Z} , while the 2x2 subblock contains all \E{a}{A} with a = x,y and A = X,Y. The Ashtekar scalar constraint H is density weight 2, and the new Lagrange multiplier for the scalar constraint is a weight -1 lapse \ut{N} which is related to \mbox{${\rm N}'$} as follows: \begin{eqnarray} \ut{N} \E{z}{Z} &=& (\rm N /\sqrt{g}) (\sqrt{g} e^z_Z) \nonumber \\ &=& \rm N \sqrt{g^{zz}} \nonumber \\ &=& \mbox{${\rm N}'$} . \label{eq:2.4} \end{eqnarray} Hence in the Ashtekar approach we will be absorbing factors of \E{z}{Z} into the densitized lapse. Since $\ut{N} H = \mbox{${\rm N}'$} (H/\E{z}{Z}) $, the new scalar constraint $H/\E{z}{Z} $ becomes a rational function of the basic fields, rather than a polynomial function. This complication is a price which must be paid in order to secure closure. We consider it a small price. Firstly, the constraints should close as a matter of principle, in order to have a consistent quantization and full four-dimensional diffeomorphism invariance. Secondly, the presence of the $1/\E{z}{Z} $ factor actually makes the Hamiltonian \underline{less} singular. $H$ contains products of several operators, all evaluated at the same point z, which can lead to undefined $\delta (z-z)$ factors when $H$ acts upon a wave functional. To avoid such factors, Husain and Smolin \cite{HSm} must regulate $H$ by point-splitting. However, most of the terms in $H$ contain a factor of $\E{z}{Z} $, which is removed by the $1/\E{z}{Z} $, leaving behind a simpler operator which cannot produce a $\delta (z-z)$ and does not have to be regulated. To study the term where the $1/\E{z}{Z} $ does not cancel, we recall that \E{z}{Z} is conjugate to the complex Ashtekar connection \A{A}{a} , \begin{equation} [\E{a}{A} ,\A{B}{b} ] = -\hbar \delta (z - z')\delta ^a_b \delta ^B_A, \label{eq:2.7} \end{equation} Therefore the action of \E{z}{Z} on a wave functional $\psi = \psi [{\rm A}]$ is $\E{z}{Z} \psi = -\hbar \delta \psi / \delta \A{Z}{z} $. We expect the \A{Z}{z} dependence of $\psi $ to be holonomic. \begin{eqnarray} \psi & =& \cdots X(z_3) \exp [i\int_{z2}^{z3} \A{Z}{z} S_Z dz]X(z_2)\exp [i\int_{z1}^{z2} \A{Z}{z} S_Z dz] \cdots \nonumber \\ &:=& \cdots X(z_3)U(z_3, z_2)X(z_2)U(z_2,z_1)\cdots . \label{eq:2.8} \end{eqnarray} The X are assumed to be operators in the Lie algebra of SU(2), and independent of \A{Z}{z} . Explicit factors of i are present, because we use the usual, Hermitian SU(2) generators $S_I$. Then the direct action of \E{z}{Z} on the wave functional is \begin{equation} \E{z}{Z} (z)\psi = \cdots + \cdots X(z_3)[\hbar \theta (z_3-z) \theta (z-z_2) U(z_3, z_2)(-iS_Z)]X(z_2)U\cdots + \cdots , \label{eq:2.8a} \end{equation} one such term for each U. To find the action of $[1/\E{z}{Z}]$, multiply both sides of equation~(\ref{eq:2.8a}) by $\theta (z_3-z)\theta (z-z_2)$, in order to project out the term exhibited explicitly on the right; then multiply both sides of equation~(\ref{eq:2.8a}) by $\hbar /\E{z}{Z}$. \begin{eqnarray} [\hbar /\E{z}{Z} (z)] &\cdot &\theta (z_3-z)\theta (z-z_2)[\cdots X(z_3)U(z_3, z_2)(-iS_Z)X(z_2) \cdots ] \nonumber \\ & = & \theta (z_3-z)\theta (z-z_2)[\cdots X(z_3)U(z_3, z_2)X(z_2)\cdots ] + \nonumber \\ & & + const. \label{eq:2.8c} \end{eqnarray} In order to make this look like $\hbar /\E{z}{Z}$ acting upon $\psi$, multiply $\psi$ by the following partition of unity: \begin{equation} 1 = [\theta (z-z_n) + \theta (z_n-z)\theta(z-z_{n-1}) + \cdots + \theta (z_3-z)\theta (z-z_2) + \cdots ]. \label{eq:2.8d} \end{equation} Therefore \begin{eqnarray} [\hbar /\E{z}{Z} (z)]\psi & = & \cdots + \cdots X(z_3)[\theta (z_3- z) \theta (z-z_2) U(z_3, z_2)(-iS_Z)^{-1}]X(z_2)U\cdots + \nonumber \\ & & + \cdots + {\rm const.}. \label{eq:2.8b} \end{eqnarray} Evidently the action of $\hbar /\E{z}{Z}$ on $\psi$ is quite mild. There is not even a $\delta (z-z')$ factor, let alone a $\delta (z - z)$ {}. If the X are helicity-changing operators, the eigenvalue of $S_Z$ in equation~(\ref{eq:2.8b}) will vary from one U to the next. If we use the 2x2 Pauli representation, the eigenvalue never vanishes and $(S_Z)^{-1}$ is always finite. However, in future work we shall use the (2j+1)x(2j+1) representation, where $S_Z$ can have a zero eigenvalue if j is integer. When $S_Z$ vanishes, so that $U(z_3,z_2)$ is unity, one may replace the square bracket in equation~(\ref{eq:2.8b}) by \begin{equation} [-\theta (z_3-z) \theta (z-z_2) U(z_3, z_2)\int_2^3 \A{Z}{z} dz]. \label{eq:2.8e} \end{equation} As a check, when equation~(\ref{eq:2.8b}) is inverted by multiplying both sides by $\E{z}{Z}/\hbar $, the square bracket~(\ref{eq:2.8e}) gives the same answer for vanishing $S_Z$ as the square bracket~(\ref{eq:2.8b}) gives for finite $S_Z$. We should also check that \E{z}{Z} is not a gauge artifact (which can be gauged to zero!). Despite its contravariant z index, \E{z}{Z} is a scalar function under diffeomorphisms, \begin{equation} \delta \E{z}{Z} = \rm N ^z \partial _z \E{z}{Z} \label{eq:2.5} \end{equation} The \E{z}{Z} field is both contravariant and weight 1, and in effect the two transformation properties cancel each other in one space dimension, leaving an ordinary scalar. Another way to see the scalar behavior is to relate \E{z}{Z} to the metric variables A, B, W, and D introduced at equation~(\ref{eq:2.1}). Using the same relationships as at equation~(\ref{eq:2.4}), we find \begin{equation} \E{z}{Z} = exp(A). \label{eq:2.6} \end{equation} Since A occurs in the x,y sector of the metric~(\ref{eq:2.1}), and this sector transforms as a scalar under diffeomorphisms, the function A is a scalar. Hence $\E{z}{Z} $ cannot be gauged away. We now pass to the details. In a coordinate system where both \E{a}{A} and \A{A}{a} fields are block diagonal, the total Hamiltonian reduces to \begin{eqnarray} H_T &=& \mbox{${\rm N}'$} [\epsilon _{MN}\E{x}{M} \E{y}{N} (\E{z}{Z} )^{-1} \epsilon _{AB} \A{A}{x} \A{B}{y} + \epsilon _{MN} \E{b}{M} {\rm F}^N_{zb}] \nonumber \\ & & + i\rm N ^z\E{b}{M} {\rm F}^M_{zb} \nonumber \\ & & -i{\rm N_G} [\partial _z\E{z}{Z} - \epsilon _{IJ}\E{a}{I} \A{J}{a} ] \nonumber \\ &\equiv & \mbox{${\rm N}'$} H_S + \rm N^zH_z + {\rm N_G}H_G, \label{eq:2.9} \end{eqnarray} where\begin{equation} {\rm F}^N_{zb} = \partial _z\A{N}{b} -\epsilon _{NQ}\A{Z}{z} \A{Q}{b}, \label{eq:2.10} \end{equation} and the ${\rm N_G}$ term (the Gauss constraint) is the generator of local SU(2) rotations around the Z axis. The lapse has been renormalized as at equation~(\ref{eq:2.4}), by removing a factor \E{z}{Z} from the scalar constraint. From now on the "scalar constraint" will mean the expression $H_S$ multiplying \mbox{${\rm N}'$}\ in equation~(\ref{eq:2.9}), namely, the usual Ashtekar scalar constraint H divided by \E{z}{Z} . We have operator-ordered equation~(\ref{eq:2.9}) in a way which anticipates the following section, where we shall consider solutions $\psi $ which depend on \A{Z}{z} and the four \Etld\ in the 2x2 sector. If we call these five commuting variables the Q variables ($\psi = \psi (Q)$) and the five conjugate variables the P variables, then we have ordered P's to the right, Q's to the left in $H$. We shall carry out the proof of closure for this specific choice of Q's and this specific ordering, but the proof would also go through for the other popular choice of Q's. in which the five A's are chosen as Q's (and the A's are ordered to the left in $H$). Now let us ask which commutators, or which parts of which commutators, are likely to give trouble. First of all, it is easy to check that the \underline{classical} commutators (or rather, the classical Poisson brackets) close on pure constraints, with no undesirable factors of $g^{zz}$ or the Ashtekar analog of $g^{zz}$. Since the quantum commutators are designed to reproduce exactly the same fields as the classical Poisson brackets, there will be no factor of $g^{zz}$ in the quantum case either. We will get the same fields; and the only thing which can go wrong is that the commutator yields a P field to the left of a Q field. Remembering that each constraint is a sum of terms of the form f(Q)g(P), we want, schematically, \begin{equation} [f_1(Q)g_1(P),f_2(Q)g_2(P)] = f_3(Q)g_3(P). \label{eq:2.11} \end{equation} There would be trouble if P's occured to the left of Q's on the right-hand side, for example $Pf_3(Q)g_3(P)$. We now show that almost all the terms in a typical constraint commutator $[H_i,H_j]$ will give no trouble. Each term in this commutator will look like the left-hand side of equation~(\ref{eq:2.11}). Write this term out using the identity \begin{eqnarray} \label{eq:2.12} [AB,CD]& =& AC[B,D]+A[B,C]D+ \nonumber \\ & & +C[A,D]B+[A,C]DB. \\ \label{eq:2.13} [f_1g_1,f_2g_2]&=& 0 + f_1[g_1,f_2]g_2 + f_2[f_1,g_2]g_1 + 0. \end{eqnarray} On the right-hand side, the f and g factors outside the commutators are in the correct order (Q's to the left). Therefore if the commutators on the right yield fields which commute among themselves, the entire expression on the right can be ordered correctly, and the term gives no trouble. In particular, if either f or g is a monomial, say $g\sim $P, then the commutator of any f with g yields only Q fields, and the term can be ordered correctly. Examination of equation~(\ref{eq:2.9}) shows that all the terms in $H_T$, except the $(\E{z}{Z} )^{-1}$ term, are monomials in the P's, of the form $Q^2P$ or $QP$; and one term is independent of the P's (pure Q) which is even better. Therefore the only commutators we have to check are the ones involving the $(\E{z}{Z} )^{-1}$ term, \begin{eqnarray} H_E &:= &\mbox{${\rm N}'$} [\epsilon _{MN}\E{x}{M} \E{y}{N} (\E{z}{Z} )^{- 1}\epsilon _{AB} \A{A}{x} \A{B}{y} \nonumber \\ &\sim &Q^2(1/P)P^2. \label{eq:2.14} \end{eqnarray} We now investigate commutators of $H_E$ with Q terms, QP terms, $Q^2P $ terms, and finally commutators of $H_E$ with itself. (a). Commutators of $H_E$ with Q and QP terms. For this case, in commutator~(\ref{eq:2.11}), \underline{both} $f_2$ and $g_2$ are monomials or constants. It follows immediately that both commutators on the right-hand side of equation~\ref{eq:2.13} involve at least one monomial and can be correctly ordered. Even if $f_2$ is \A{Z}{z} , there is no problem, since the commutator with (\E{z}{Z} )$^{-1}$, \begin{equation} [(\E{z}{Z} )^{-1},\A{Z}{z} ] = \hbar \delta (z-z') (\E{z}{Z} )^{-2}, \label{eq:2.14a} \end{equation} yields factors which commute among themselves. Equation~(\ref{eq:2.14a}) may be proven by multiplying it on left and right by \E{z}{Z} . (b). Commutators of $H_E$ with $Q^2P$ terms. These are terms of the form \E{b}{M} \A{Z}{z} \A{Q}{b} coming from the F's (field strengths) in the scalar and vector constraints, equations {}~(\ref{eq:2.9})-(\ref{eq:2.10}). (b1). When commuting the scalar constraint with itself, for example, we get commutators of the form \begin{eqnarray*} [Q_E^2(1/P_E)P_E^2(z),Q^2P(z')] &+& [Q^2P(z),Q_E^2(1/P_E)P_E^2(z')] \\ &=&\cdots + Q_E^2[(1/P_E)P_E^2(z),Q^2(z')]P + \\ & & +Q_E^2[Q^2(z),(1/P_E)P_E^2(z')]P, \end{eqnarray*} where subscripts E denote fields coming from $H_E$, and $\cdots $ indicates harmless commutators involving the monomial P. The two commutators on the last line differ by a minus sign and an interchange of z and $z'$. The interchange affects nothing, since the commutator is proportional to $\delta (z-z')$. Therefore the two commutators cancel each other, and the term is harmless. (b2). When commuting the scalar constraint with the vector constraint, we get \begin{eqnarray*} [Q^2P(z),Q^2(1/P)P^2(z')]&=& \cdots + Q_2(z')[Q^2(z),(1/P)P^2(z')]P(z) \\ &=& \cdots + \mbox{$^{(2)}\!\tilde{\rm E} $} [\E{b}{M} \A{Z}{z} ,(\E{z}{Z} )^{- 1}\epsilon _{AB} \A{A}{x} \A{B}{y}]i\A{N}{b} \epsilon _{MN} \\ &=& \cdots + \mbox{$^{(2)}\!\tilde{\rm E} $} \E{b}{M} [\A{Z}{z} ,(\E{z}{Z} )^{- 1}]\epsilon _{AB} \A{A}{x} \A{B}{y} i\A{N}{b} \epsilon _{MN} + \\ & & + \mbox{$^{(2)}\!\tilde{\rm E} $} (\E{z}{Z} )^{-1}[\E{b}{M} ,\epsilon _{AB} \A{A}{x} \A{B}{y}]\A{Z}{z} i\A{N}{b} \epsilon _{MN} \\ &=& \cdots - \mbox{$^{(2)}\!\tilde{\rm E} $} \E{b}{M} [\hbar \delta (z-z')(\E{z}{Z} )^{-2}]\epsilon _{AB} \A{A}{x} \A{B}{y} i\A{N}{b} \epsilon _{MN} +0. \end{eqnarray*} On the third line we have used the ABCD identity, equation~(\ref{eq:2.12}). On the last line, we can commute the i\A{N}{b} factor to the left, until it forms the expression $i\E{b}{M} \A{N}{b} \epsilon _{MN}$. We can subtract from this expression the Gauss constraint $H_G$, equation~(\ref{eq:2.9}). This causes no problems, since the Gauss constraint commutes with everything to the right of $i\E{b}{M} \A{N}{b} \epsilon _{MN}$, hence can be commuted to the far right where it will eventually annihilate the wave functional. Since $i\E{b}{M} \A{N}{b} \epsilon _{MN} - H_G = i\partial _z\E{z}{Z} $, the last line becomes \begin{displaymath} \cdots +i\mbox{$^{(2)}\!\tilde{\rm E} $} [\hbar \delta (z-z')\partial _z(\E{z}{Z} )^{- 1}]\epsilon _{AB} \A{A}{x} \A{B}{y}. \end{displaymath} This now has the same form as the corresponding term in the classical calculation, and moreover the operators are correctly ordered (Q's to the left). The calculation just completed, however, suggests a new way in which the constraints might fail to close. Suppose that at some point in the calculation it is necessary to insert the Gauss constraint in the middle of a term (as was done just above); if the $H_G$ factor cannot be commuted to the far right, then closure will be spoiled. Fortunately, the classical calculation once again comes to the rescue. Since the classical calculation has the same pattern of fields as the quantum calculation, a Gauss insertion is necessary in the quantum calculation if and only if a Gauss insertion is necessary at the same point in the classical calculation. It turns out that the \underline{only} point where a Gauss insertion occurs, classically, is in the [vector,scalar] commutator term just considered, and this insertion is harmless. (c). Finally, we consider the commutator of $H_E$ with itself. Using the ABCD identity~(\ref{eq:2.12}), we get \[ [H_E(z),H_E(z')] = \mbox{$^{(2)}\!\tilde{\rm E} $} (z)(\E{z}{Z} )^{-1}[\epsilon _{AB} \A{A}{x} \A{B}{y} (z),\mbox{$^{(2)}\!\tilde{\rm E} $} (z')(\E{z}{Z} )^{-1}]\epsilon _{CD} \A{C}{x} \A{D}{y} (z')+ \] \[ \mbox{$^{(2)}\!\tilde{\rm E} $} (z')(\E{z}{Z} )^{-1}[\mbox{$^{(2)}\!\tilde{\rm E} $} (z)(\E{z}{Z} )^{-1},\epsilon _{CD} \A{C}{x} \A{D}{y} (z')]\epsilon _{AB} \A{A}{x} \A{B}{y} (z). \] The two commutators cancel, after a relabeling AB$\leftrightarrow $CD in the second commutator. This completes the proof that the constraints close. It is also easily verified that the constraint $H_z$ generates diffeomorphisms in the z direction. (In fact $H_z$ fails to generate diffeomorphisms only when the P's are ordered to the left \cite{knotsol}.) $H_z$ is not quite the diffeomorphism generator, but differs from that generator by a term of the form $N^z\A{Z}{z} H_G$, $H_G$ the Gauss constraint. A term of this form can be added to $H_z$ without affecting closure, since the added term is linear in P and therefore harmless. Since there is no factor of $g^{zz}$ or other fields in the constraint algebra, it is a true Lie algebra (structure ``constants'' are at most $\delta $ functions, not functions of fields). The structure of this Lie algebra is very simple. It breaks up into two commuting subalgebras generated by $(H_z \pm H_S)/2$, \begin{equation} \int dz\int dz'[M(z)(H_z \pm H_S)/2,N(z')(H_z \pm H_S)/2] \\ = i\hbar \int dz(M\partial _zN -N\partial _zM)(H_z \pm H_S)/2. \label{eq:2.15} \end{equation} Presumably these generators may be interpreted physically as displacements along the light cone, in directions (z $\pm $ct). \section{The Reality Constraints} Since the Ashtekar connections are complex, they obey reality constraints of the form A + A* = 2 Re A = known function of the \Etld. For completeness, and to establish certain detailed formulas which we will need later, we sketch a derivation of these constraints. The derivation ends at equation~(\ref{eq:5.12}). At equation~(\ref{eq:5.13}) we propose a measure to enforce these constraints. We start from the four-dimensional connection \begin{equation} 2{\rm G}\, ^{(4)}\!\A{IJ}{a} = \omega ^{IJ}_a +i\epsilon ^{IJ}_{MN}\omega ^{MN}_a, \label{eq:5.1} \end{equation} where G is the Newtonian constant and $\omega $ is the SL(2,C) Lorentz connection. After the 3 + 1 splitup \cite{Alect, Rovlect}, one obtains the SU(2) connection which is canonically conjugate to $\E{a}{S}$ (equation~(\ref{eq:1.2})). \begin{eqnarray} 2{\rm G}\A{S}{a} &\:= & \epsilon _{MSN}\, ^{(4)}\!\A{MN}{a} \nonumber \\ &=&\epsilon _{MSN}\omega ^{MN}_a - 2i\omega ^{TS}_a. \label{eq:5.2} \end{eqnarray} {}From this, the real part of A is \begin{equation} {\rm G}[\A{S}{a} + \A{S}{a} *] = \epsilon _{MSN}\omega ^{MN}_a , \label{eq:5.3} \end{equation} or when these equations are written out for the 1x1 and 2x2 subblocks, \begin{eqnarray} {\rm G}[\A{Z}{z} + \A{Z}{z} *] &=&-2\omega ^{XY}_z; \\ \label{eq:5.3a} {\rm G}[\A{I}{i} + \A{I}{i} *]&=&2\epsilon _{IJ} \omega ^{ZJ}_i . \label{eq:5.3b} \end{eqnarray} These ``reality constraints'' relate the A's to the \Etld 's , since $\omega = \omega (\tilde {\rm E} )$. The next step is to exhibit this $\tilde {\rm E}$ dependence. This is done by first relating the $\omega$ 's to the triads $e^I_i $ and inverse triads $e^i_I$, then relating the $ e^I_i$ and $ e^i_I$ to the $ \tilde {\rm E}$ . The requirement that the triads have zero covariant derivative leads to \begin{equation} \omega ^{IJ}_a = [\Omega _{i[ja]} + \Omega _{j[ai]} - \Omega _{a[ij]}]e^{iI} e^{jJ}, \label{eq:5.4} \end{equation} where \begin{equation} \Omega _{i[ja]} = e_{iM}[\partial _je^M_a - \partial _ae^M_j]/2. \label{eq:5.5} \end{equation} In the present case the triad matrix is block diagonal, with 2x2 and 1x1 subblocks, and these equations simplify considerably. Also, from equations ~(\ref{eq:5.3a}) and ~(\ref{eq:5.3b}) we shall need only \begin{eqnarray} \omega^ {XY}_z&=& [e^X_i\partial _ze^{Yi} - e^Y_i\partial _ze^{Xi}]; \label{eq:5.6} \\ \omega ^{ZJ}_i &=& -\partial _zg_{ij}e^{zZ}e^{jJ}/2, \label{eq:5.7} \end{eqnarray} where the indices i,j range over x,y only. Now we replace metric, triads, and inverse triads by \Etld\ fields. The inverse triad fields are easiest to replace. From equation~(\ref{eq:1.1}) \begin{eqnarray} e^a_A &=& \E{a}{A} /\sqrt g \nonumber \\ &=&\E{a}{A} /\sqrt{\E{z}{Z}\mbox{$^{(2)}\!\tilde{\rm E} $} }. \label{eq:5.8} \end{eqnarray} For the metric and triad fields, the strategy is to express them in terms of the inverse triads, then replace the latter. For example in the 2x2 subblock, \begin{eqnarray} ^{(2)}g_{ab}& =& \epsilon _{am}\epsilon _{bn}{}^{(2)}\!g^{mn}\, ^{(2)}\!g\nonumber \\ &=&\epsilon _{am}\epsilon _{bn} \E{m}{M} \E{n}{M} ^{(2)}\!g/g \nonumber \\ &=&\epsilon _{am}\epsilon _{bn} \E{m}{M} \E{n}{M} \E{z}{Z} / \mbox{$^{(2)}\!\tilde{\rm E} $}; \label{eq:5.9} \end{eqnarray} \begin{eqnarray} e^M_m &=& \epsilon _{MN}\epsilon _{mn}e^n_N\, ^{(2)}e \nonumber \\ &=& \epsilon _{MN}\epsilon _{mn} \E{n}{N} \sqrt{\E{z}{Z} /\mbox{$^{(2)}\!\tilde{\rm E} $} } . \label{eq:5.10} \end{eqnarray} We make these replacements in equations~(\ref{eq:5.6}) and {}~(\ref{eq:5.7}) and obtain \begin{eqnarray} \omega^ {XY}_z&=&-[\epsilon _{mn} \E{m}{M} \partial _z\E{n}{M}]/2 \mbox{$^{(2)}\!\tilde{\rm E} $} ; \\ \label{eq:5.11} \omega ^{ZJ}_i &=& -[\E{j}{J} /2\mbox{$^{(2)}\!\tilde{\rm E} $} ]\partial _z[\epsilon _{im}\epsilon _{jn} \E{m}{M} \E{n}{M} \E{z}{Z} /\mbox{$^{(2)}\!\tilde{\rm E} $} ]. \label{eq:5.12} \end{eqnarray} After these algebraic preliminaries, we are ready to consider the measure. Since our complete set of commuting observables are the four \Etld\ in the 2x2 X,Y sector, plus the complex connection \A{Z}{z}, we try a dot product of the form \begin{equation} <\phi \mid \psi > = \int \mbox{$\phi\!*$}\psi \mu d^4\tilde {\rm E} d^2A, \label{eq:5.13} \end{equation} where $d^2A \equiv dRe\A{Z}{z} dIm\A{Z}{z} $. The measure $\mu $ must satisfy several requirements. (i) It must guarantee the quantum form of the reality constraints. \begin{equation} <\phi \mid A\psi > + <A\phi\mid \psi > = 2<\phi\mid {\rm ReA}\psi >. \label{eq:5.14} \end{equation} (ii) It must guarantee the invariance of $\mu d^4\tilde {\rm E} d^2A$ under transformations generated by the scalar, vector, and Gauss constraints. (iii) It must contain enough gauge-fixing delta functions to remove the usual unbounded integrations over infinite numbers of gauge copies. Note that (ii) requires only invariance under the constraints, not invariance under four- dimensional diffeomorphisms. In a 3 + 1 formalism, one does not have the proper set of fields to implement the latter invariance, essentially because all fields are evaluated on a constant time hyperslice, whereas four-dimensional diffeomorphisms move fields off the hyperslice \cite{IK}. As yet we do not know how to satisfy requirements (ii)- (iii). We shall, however, propose a $\mu $ which will satisfy requirement (i). We set \begin{equation} \mu = \delta [{\rm G}(\A{Z}{z} + \A{Z}{z}*) + 2\omega ^{XY}_z]. \label{eq:5.15} \end{equation} {}From equation~(\ref{eq:5.3a}), this delta function enforces the \A{Z}{z} reality constraint. The surprising fact is that it also enforces the remaining reality constraints~(\ref{eq:5.3b}) as well, as we shall prove now. First we should clarify our notation. Since our integration variables are Re\A{Z}{z} and Im\A{Z}{z} , not A and A*, the \A{Z}{z} functional derivative really means \begin{equation} \delta /\delta \A{Z}{z} := [\delta /\delta Re\A{Z}{z} + (1/i)\delta /\delta Im\A{Z}{z} ]/2, \label{eq:5.16} \end{equation} which follows from ReA = (A + A*)/2, etc. We can then check that, since the $\phi$ in the bra, equation~(\ref{eq:5.14}) , is complex conjugated, \begin{eqnarray} \delta \mbox{$\phi\!*$} /\delta \A{Z}{z} &=& \delta \mbox{$\phi\!*$}[{\rm A*}]/ \delta {\rm A} \nonumber \\ &=&\int dz'[ \delta \mbox{$\phi\!*$}/\delta \A{Z}{z} (z')\!*][\delta \A{Z}{z}\! */\delta Re\A{Z}{z} + (1/i)\delta \A{Z}{z}\! */\delta Im\A{Z}{z} ]/2 \nonumber \\ &=& 0, \label{eq:5.17} \end{eqnarray} as expected. Now write \begin{eqnarray} <\phi \mid \A{I}{i} (z) \psi >& =& -\hbar \int \mbox{$\phi\!*$}\mu \delta \psi /\delta \E{i}{I} (z) \nonumber \\ &=&+\hbar \int [\delta \mbox{$\phi\!*$}/\delta \E{i}{I} \mu \psi + \mbox{$\phi\!*$}\delta \mu /\delta \E{i}{I} \psi]\nonumber \\ &=&-<\A{I}{i} \phi \mid \psi > + \hbar \int \mbox{$\phi\!*$} \delta \delta [{\rm G(A + A*)} + 2\omega ^{XY}_z]/\delta \A{Z}{z} (z') \times \nonumber \\ & & \times [2\delta \omega ^{XY}_z(z')/\delta \E{i}{I} (z)] dz'\psi \nonumber \\ &=&-<\A{I}{i} \phi \mid \psi > - (\hbar /{\rm G})\int \mbox{$\phi\!*$} \mu [2\delta \omega ^{XY}_z(z')/\delta \E{i}{I} (z)] dz' \delta \psi /\delta \A{Z}{z} (z') \nonumber \\ &=&-<\A{I}{i} \phi \mid \psi > - (1/{\rm G})\int \mbox{$\phi\!*$} \mu [2\delta \omega ^{XY}_z(z')/\delta \E{i}{I} (z)] dz'\E{z}{Z} (z') \psi . \label{eq:5.18} \end{eqnarray} Again, the $\delta /\delta \A{Z}{z}$ is really a sum of ReA and ImA functional derivatives, as at equation~(\ref{eq:5.16}), with $ \delta \mu /\delta Im\A{Z}{z} $ vanishing. The last square bracket can be rewritten using equation~(\ref{eq:5.11}) and $^{(2)} \tilde {\rm E} = \epsilon _{ij}\epsilon _{IJ}\E{i}{I} \E{j}{J}/2$. \begin{eqnarray} -\E{z}{Z} (z')2\delta \omega ^{XY}_z(z')/\delta \E{i}{I} (z) &=& \E{z}{Z} [(\delta \epsilon _{in}\partial _{z'}\E{n}{I} + \epsilon _{mi}\E{m}{I} \partial _{z'}\delta )/\mbox{$^{(2)}\!\tilde{\rm E} $} \nonumber \\ & & - \epsilon _{mn} \E{m}{M} \partial _{z'}\E{n}{M} \delta \epsilon _{ij}\epsilon _{IJ}\E{j}{J} /(\mbox{$^{(2)}\!\tilde{\rm E} $} )^2], \label{eq:5.19} \end{eqnarray} where $\delta = \delta (z - z')$. Now use $\epsilon _{mn} \epsilon _{ij} = \epsilon _{mi} \epsilon _{nj} +\epsilon _{mj} \epsilon _{in}$ in the last term of equation~(\ref{eq:5.19}). The two terms involving $ \epsilon _{in}$ cancel. After integrating by parts to remove the derivative from the delta function in the second term, one may replace the $\E{m}{I}$ by $\E{m}{M}\delta^M_I = \E{m}{M} [\epsilon _{nj}\epsilon _{IJ}\E{n}{M} \E{j}{J} /\mbox{$^{(2)}\!\tilde{\rm E} $} ]$. We then have \begin{eqnarray} -\E{z}{Z}2 \delta \omega ^{XY}_z(z')/\delta \E{i}{I} &=&-\partial _{z'}[\E{z}{Z} \epsilon _{mi}\epsilon _{nj}\E{m}{M} \E{n}{M} /\mbox{$^{(2)}\!\tilde{\rm E} $} ]\delta (z - z')\epsilon _{IJ}\E{j}{J} /\mbox{$^{(2)}\!\tilde{\rm E} $} \nonumber \\ &=& 2\omega ^{ZJ}_i \epsilon _{IJ}\delta (z - z') \nonumber \\ &=& 2G {\rm Re}\A{I}{i} \delta (z - z'), \label{eq:5.20} \end{eqnarray} where the third line uses equation~(\ref{eq:5.12}) and the last line uses equation~(\ref{eq:5.3b}). Inserting the result~(\ref{eq:5.20}) into equation~(\ref{eq:5.18}), we obtain the reality condition for the \A{I}{i} field, QED. \section{Directions for Further Research.} For the plane wave problem, we have constructed a constraint algebra which closes after a simple renormalization of the lapse function. We have argued that the cost of this renormalization (rational, rather than polynomial constraints) is small compared to the benefits (consistent constraints in the quantum-mechanical theory; full diffeomorphism invariance in the classical theory). We have also made modest progress toward constructing a measure. It is a standard result that the Hamiltonian has surface terms whenever the spatial manifold is non-compact \cite{deW}. In future work, we intend to describe these surface terms. They are surprising: in the plane wave case, it is not automatically true that the Gauss constraint term in the Hamiltonian falls to zero at infinity. Because of these additional terms at infinity, one must exercise care when interpreting any wavefunctional involving holonomies defined over open contours. In the planar case a holonomy $\exp (i\int \A{Z}{z}S_z)$ integrated over a closed contour is necessarily zero, since the z contour must retrace itself. Therefore it is natural to consider open contours and extend the endpoints to $z = \pm \infty$, to respect spatial diffeomorphism invariance. It is possible to enrich this elemental holonomic structure in two ways. First, insert \Etld\ operators at various points along the holonomy; the wavefunctional then looks like a Rovelli-Smolin $T^n$ operator \cite{RovSmo} (defined over an open rather than closed contour). Second, replace the usual 2x2 $S_z$ matrices in the holonomy and in \Etld\ by the (2j + 1) x (2j + 1) spin-j representation. The resultant structure is reminiscent of a symmetric state, or spin network state \cite{symmsta} , with the holonomies corresponding to flux lines of spin j and the \Etld\ operators corresponding to vertices. For appropriate choice of the \Etld, a wavefunctional constructed in this manner is annihilated by the Hamiltonian at {\it finite} values of z. In three spatial dimensions, this would be essentially the end of the story: the flux exiting at infinity is irrelevant, since Gauss rotations at infinity are not allowed. The surviving surface terms in the Hamiltonian simply give the ADM energy. In one spatial dimension, however, the Gauss term contributes at infinity. One could add Fermionic matter to the theory \cite{Romano, Morales} and terminate the flux lines on Fermions at $\pm \infty$. However, adding Fermions probably complicates the theory unnecessarily. In one-dimensional QED, for example, one can learn quite a bit by studying electromagnetic plane waves at finite z, while ignoring what happens to the wave at infinity. In a future publication we will adopt this philosophy and study the finite z properties of the open flux line solutions. Work is also in progress on solutions involving closed flux lines.
\section{Background} \label{sec:Background} A current trend in logic is to attempt to incorporate semantic information into the domain of deduction, \cite{gabbay94}, \cite{barwiseetal95}. An area for which this strategy is particularly useful is the problem of categorial grammar parsing. The categorial grammar research programme requires the use of a range of logical calculi for linguistic description. Some researchers have considered labelled deduction as a tool for implementing categorial parsers \cite{moortgat92}, \cite{morrill95}, and this paper can be seen as a new contribution to this field. In this paper we aim for a modular approach, in which the basic grammar is kept constant, while different calculi can be implemented and experimented with by constraining the derivations produced by the theorem prover. At present, our system covers the classical Lambek Calculus, \calcL{}, as well as the non-associative Lambek calculus \calcNL{}, \cite{lambek61}, and variants such as Van Benthem's \cite{vanbenthem88} \calcLP{}, \calcLPC{}, \calcLPE{} and \calcLPCE{}, and their non-associative counterparts. The system is based on labelled analytic deduction, particularly on the LKE method, developed by D'Agostino and Gabbay \cite{dagostino94}. LKE is similar to a Smullyan-style tableau system, in which the derivations obey the sub-formula principle, but it improves on efficiency by restricting the number of branching rules to just one. Different categorial logics are handled by assigning different properties to the labelling algebra, while the basic syntactic apparatus remains the same. This allows the user to experiment with various linguistic properties without having in principle to modify the grammar itself. The basic structure of the paper is as follows. In section \ref{sec:Family}, we introduce the family of categorial calculi, and discuss some of the linguistic arguments which have been put forward in the literature with regard to these calculi. In section \ref{sec:Framework}, we introduce the logical apparatus on which the system is based, describe the algorithm and prove some of the properties mentioned in section \ref{sec:Family} within this framework. We also show how different grammars can be characterised and present a worked example. In section \ref{sec:Comparison} the system is compared with other strategies for dealing with multiple categorial logics, such as hybrid formalisms and unification-based Gentzen-style deduction. In this section, we also suggest some ways to improve the efficiency of the system, and strategies for dealing with the complexity of labelled unification. \subsection{A Family of Categorial Calculi and Their Linguistic Applications} \label{sec:Family} Categorial Grammars can be formalised in terms of a hierarchy of well understood and mathematically transparent logics, which yield as theorems a range of combinatorial operations. However the precise nature of the combinatorial power required for an adequate characterisation of natural language is still very much a matter of debate. For this reason, it is desirable to have a means of systematically testing the linguistic consequences of adopting various calculi. In this section we give an overview of the linguistic applications of some of the calculi in the hierarchy, with a view towards motivating the usefulness of a generic categorial theorem prover as a tool for linguistic study. The combinatorial possibilities of expressions in general can be characterised in terms of {\em reduction laws}. In {\em R1}-{\em R6} below, we give some reduction laws discussed in \cite{moortgat88}, which have been found to be linguistically useful.\footnote{We adopt the Lambek notation, in which X$/$Y is a function which ``takes'' a Y to its right to yield an X, and Y$\backslash$X is a function which ``takes'' a Y to its left to yield an X.}. {\small \begin{center} \begin{tabular}[ht]{llllll} {\em R1:} & {\em Application} & {\em R2:} & {\em Composition} & {\em R3:} & {\em Associativity} \\ & X$/$Y, Y \EN X & & X$/$Y, Y$/$Z \EN X$/$Z & & (Z$\backslash$X)$/$Y \EN Z$\backslash$(X$/$Y) \\ & Y, Y$\backslash$X \EN X & & Z$\backslash$Y, Y$\backslash$X \EN Z$\backslash$X & & Z$\backslash$(X$/$Y) \EN (Z$\backslash$X)$/$Y \\ & & & & & \\ {\em R4:} & {\em Lifting} & {\em R5:} & {\em Division (main functor)} & {\em R6:} & {\em Division (sub. functor)} \\ & X \EN Y$/$(X$\backslash$Y) & & X$/$Y \EN (X$/$Z)$/$(Y$/$Z) & & X$/$Y \EN (Z$/$X)$\backslash$(Z$/$Y)\\ & X \EN (Y$/$X$)\backslash$Y & & Y$\backslash$X \EN (Z$\backslash$Y)$\backslash$(Z$\backslash$X) & & Y$\backslash$X \EN (Y$\backslash$Z)$/$(X$\backslash$Z)\\ \end{tabular} \end{center}} It is possible to define a hierarchy of logical calculi, each of which admits one or more of {\em R1}-{\em R6} as theorems; from the purely applicative calculus {\sf AB}, of Ajdukiewicz and Bar-Hillel, \cite{ajduk35}, which supports only {\em R1}, to the full Lambek calculus {\sf L}, which supports all the above laws. Calculi intermediate in power between {\sf AB} and {\sf L} have been explored (e.g. Dependency Categorial Grammar \cite{pickering93}), as well as stronger calculi which extend the power of {\sf L} through the addition of structural rules. Much of the interest in using categorial grammars for linguistic research derives from the possibilities they offer for characterizing a flexible notion of constituency. This has been found particularly useful in the development of theories of coordination, and incremental interpretation. For example, assuming standard lexical type assignments, the following right node raised sentence cannot be derived in {\sf AB}, but does receive a derivation in a system which includes {\em R3}, with each conjunct assigned the type indicated. \begin{examples} \item \label{ex:rnr} {} [John resents $_{S/NP}$] and [Peter envies $_{S/NP}$] Mary \end{examples} A calculus which includes composition, {\em R2}, will allow a function to apply to an unsaturated argument, and it is this property which allows Ades and Steedman \cite{steedman82} to treat long distance dependencies, and motivates much of Steedman's later work on incremental interpretation. Dowty \cite{dowty88} uses the combination of composition, {\em R2} and lifting, {\em R4}, to derive examples of non-constituent coordination such as {\em John gave mary a book and Susan a record}. We can increase the power of {\sf L} by adding the structural transformations {\em Permutation}, {\em Contraction} and {\em Expansion}, to derive the calculi {\sf LP}, {\sf LPC}, {\sf LPE} and {\sf LPCE}. The structural transformation {\em Permutation}, which removes the restrictions on the linear order of types, allows us to go beyond the purely concatenative derivations of {\sf L}. This allows us to deal with sentences exhibiting non-standard constituent order. For example, Moortgat suggests using permutation for dealing with {\em heavy NP-shift} in examples similar to the following \cite{moortgat88}: \begin{examples} \item \label{ex:comics} John gave [to his nephew $_{PP}$] [all the old comic books which he'd collected in his troubled adolescence $_{NP}$]. \end{examples} In (\ref{ex:comics}), the bracketed constituents can be ``rearranged'' via permutation so that a derivation is possible that employs the standard type ((NP$\backslash$ S)/PP)/NP for the ditransitive verb {\em gave}. In {\sf L}, while it is possible to specify a type missing an argument on its left or right periphery, it is not possible to specify a type missing an argument ``somewhere in the middle'', making it impossible to deal with non-peripheral extraction. However, as Morrill et al show, permutation provides the additional power necessary to account for this phenomenon \cite{morrilletal90}. In addition to permutation, there are also linguistic examples which motivate contraction (e.g. gapping, \cite{moortgat88}) and expansion (e.g. right dislocation, \cite{moortgat88}). However it is universally recognized that a system employing the unrestricted use of structural transformations would be far too powerful for any useful linguistic application, since it would allow arbitrary word order variation, copying and deletion. For this reason, a goal of current research is to build a system in which the resource freedom of the more powerful calculi can be exploited when required, while the basic resource sensitivity of {\sf L} is retained in the general case. One such approach is to employ structural modalities \cite{morrilletal90}, which are operators that explicitly mark those types which are permitted to be manipulated by specific structural transformations.\footnote{Systems which allow the selective use of structural transformations may be implemented in the general framework presented here, although we do not address this issue.} \section{A framework for Categorial Deduction} \label{sec:Framework} In this section we describe the theorem proving framework for categorial deduction. We start by setting up basic ideas of categorial logic, giving formal definitions of the core logical language. Then we move on to the theorem proving strategy, introducing the LKE approach \cite{dagostino94} and the algebraic apparatus used to characterise different calculi. \subsection{The core syntax} We assume that there is a finite set of atomic grammatical categories which will be represented by special symbols: NP for noun phrases, S for sentences, etc. So, the set of well-formed categories can be defined as below. \DEF{The set of well-formed categories, \cC is the smallest set which contains every basic category and which is closed under the following rule: \begin{itemize} \item[(i)] If X \EL \cC and Y \EL \cC, then X/Y, X\BS Y and X \BU Y \EL \cC \end{itemize} }{def:wff} Our purpose in this section is to define a prcedure which will enable us to verify, given an entailment relation \EN, whether or not such a relation holds for the logic being considered. Many proof procedures for classical logic have been proposed: natural deduction, Gentzen's sequents, analytic (Smullyan style) tableaux, etc. Among these, methods which conform to the sub-formula principle are particularly interesting, as far as automation is concerned. See \cite{fitting90a} for a survey. Most of these methods, along with proof methods developed for resource logics, such as Girard's proof nets (a variant of Bibel's connection method), can be used for categorial logic. Leslie \cite{leslie90} presents and compares some categorial versions of these procedures for the standard Lambek calculus L, taking into account complexity and proof presentation issues. Although tableau systems are not discussed in \cite{leslie90}, a close relative, the cut-free sequent calculus is presented as being the one which represents the best compromise between implementability and display of the proof. Smullyan style tableau systems, however, have been shown to be inherently inefficient \cite{dagostino94a}. They cannot even simulate truth-tables in polynomial time. The main reason for this is the fact that many of the Smullyan tableau expansion rules cause the proof tree to branch, thus increasing the complexity of the search. Moreover, keeping track of the structure of the derivations represents an extra source of complexity, which in most categorial parsers \cite{moortgat92,morrill95} is reflected in expensive unification algorithms employed for dealing with substructural implication. In order to cope with efficiency and generality, we have chosen the LKE system \cite{dagostino94} as the proof theoretic basis of our approach\footnote{For standard propositional logic, it has been shown that LKE can simulate standard tableau in polynomial time, but the converse is not true.}. LKE is an analytic (its derivations exhibit the sub-formula property) method of proof by refutation\footnote{A formula is proved by building a counter-model for its negation.} which has only one branching rule. In addition, its formulae are labelled according to a labelling algebra which will determine the closure conditions for the proof trees\footnote{See section \ref{sec:Comparison} for a discussion of how LKE, unlike proof nets or standard tableaux, enables us to reduce the computational cost of label unification.}. In what follows, we shall concentrate on explaining our version of the system, the heuristics that we have found useful for dealing with particularities of the calculi covered, and the relevant results for these calculi. The usual completeness and soundness results (with respect to the algebraic semantic provided) are already given in \cite{dagostino94}, so we will not discuss them here. We have mentioned that the condition for a branch to be considered closed in a standard tableau is that both a formula and its negation occur on it. The calculus defined above presents no negation, though. So, we have to appeal to some extrinsic mechanism to express contradiction. In Smullyan's original formulation, the formulae occurring in a derivation were all preceded by {\em signs}: T or F. For instance, assume that we want to prove A \IM A in classical logic. We start by saying that the formula is false, prefixing it by F, and try to find a refutation for F A \IM A. For this to be the case both T A (the antecedent) and F A (the consequent) have to be the case, yielding a contradiction. In classical logic we can interpret T and F as assertion and denial respectively, and so we can incorporate F into the language as negation, obtaining uniform notation by eliminating the need for signed formulae. In our approach, since negation is not defined in the language, we shall make use of signed formulae as proof theoretic devices. T and F will be used to indicate whether or not a certain string available for combination to produce a new one. \subsection{The generalised parsing strategy} If we had restricted the system to dealing with signed formulae, we would have a proof procedure for an implicational fragment of standard propositional logic enriched with backwards implication and conjunction. However, we have seen that the Lambek calculus does not exhibit any of the structural properties of standard logic, and that different calculi may be obtained by varying structural transformations. Therefore, we need a mechanism for keeping track of the structure of our proofs. This mechanism is provided by labelling each formula in the derivation with {\em information tokens}\footnote{See \cite{gabbay94} for a proof theoretic motivated, LDS approach, and \cite{barwiseetal95} for an approach based on a finer-grained, semantically motivated information structure.}. Labels will act not only as mechanisms for encoding the structure of the proof, from a proof-theoretic perspective, but will also serve as means to propagate semantic information through the derivation. A label can be seen as an information token supporting the information conveyed by the signalled formula it labels. Tokens may convey different degrees of informativeness, so we shall assume that they are ordered by an anti-symmetric, reflexive and transitive relation, \PO, so that an expression like x \PO y asserts that y is at least as informative as x (i.e. verifies at least as many sentences as x). We also assume that this semantic relation, ``verifies'', is closed under deductibility. It is natural to suppose that, as well as categories, information tokens can be composed. We have seen that a type S/NP can combine with a type NP to produce an S. If we assume that there are tokens x and y verifying respectively S/NP and NP, how would we represent the token that verifies S? Firstly, we define a token composition operation \CI. Then, we assume that, a priori, the order in which the categories appear in the string matters. So, a minimal information token verifying S would be x \CI y. As we shall see below, the constraints we impose on \CI will ultimately determine which inferences will be valid. For instance, if we assume that the order in which the types occur is not relevant, then we may allow permutation on the operands, so that x \CI y \PO y \CI x; if we assume that contraction is a structural property of the calculus then the string [S/NP, NP, NP] will also yield an S, since y \CI y \PO y, etc. Let's formalise these notions by defining an algebraic structure, called {\em Information frame}. \DEF{An Information Frame is a structures \cL = \OA\cP,\CI,\ID{},\PO\CA, where (i) \cP is a non-empty set of information tokens; (ii) \cP is a complete lattice under \PO; (iii) \CI is an order-preserving, binary operation on \cP which satisfies continuity, i.e., for {\em every} directed family \{z\emth{_{i}}\}, \emth{\bigsqcup} \{z\emth{_{i}} \CI x\} = \emth{\bigsqcup} \{z\emth{_{i}}\} \CI x {\em and} \emth{\bigsqcup} \{x \CI z\emth{_{i}}\} = x \CI \emth{\bigsqcup}\{z\emth{_{i}}\}; and (iv) \ID{} is an identity element in \cP. }{def:InformationFrame} Combinations of types are accounted for in the labelling algebra by the composition operator. Now, we need to define an algebraic counterpart for syntactic composition, \BU, itself. When a formula like S/NP \BU NP is verified by a token x, this is because its components were available for combination, and consequently were verified by some other tokens. Now, suppose S/NP was verified by a token, say a. What would be the appropriate token for NP, such that S/NP combined with NP would be verified by x? It certainly would not be more informative than x. Moreover, if the expression S/NP \BU NP were to stand for the composition of the (informational) meanings of its components, then the label for NP would have to verify, when combined with a, at most as much information as x. In order to express this, we define the label for NP as being {\em the greatest y s.t. x is at least as informative as a combined with y}. This token will be represented by x \opRD a. In general, x \opRD y \ED \emth{\bigsqcup} \{z\emth{\mid} y \CI z \PO x\}. An analogous operation, \opLD, can be defined to cope with cases in which it is necessary to find the appropriate label for the first operand by reversing the order of the tokens in the definition above. Some properties of \opRD \cite{dagostino94}: \begin{minipage}[b]{.48\linewidth} \FOR{Property1}{y \CI (x \opRD y) & \PO & x} \end{minipage}\hfill \begin{minipage}[b]{.48\linewidth} \FOR{Property2}{\ID{} & \PO & x \opRD x} \end{minipage} \begin{minipage}[b]{.48\linewidth} \FOR{Property3}{(x \opRD y) \CI z & \PO & (x \CI z) \opRD y} \end{minipage}\hfill \begin{minipage}[b]{.48\linewidth} \FOR{Property4}{(x \opRD y) \opRD z & \PO & x \opRD (y \CI z) } \end{minipage} \\ Having set the basic elements of our proof-theoretic apparatus, we are now able to define the components of a derivation as follows: \DEF{Signed labelled formulae (SLF) are expressions of the form S Cat : L, where S \EL \{T,F\}, Cat \EL \cC and L \EL \cL} {def:SignedFormula} A derivation, or proof will be a tree structure built according to certain syntactic rules. These rules will be called {\em expansion} rules, since their application will invariably expand the tree structure. There are three sorts of expansion rules: those which expand the tree by generating two formulae from a single one occurring previously in the derivation, those which expand the tree by combining two formulae into a third one which is then added to the tree, and the branching rule. The first kind of rule corresponds to what is called \ALPHA{}-rule in Smullyan tableaux; these rules will be called \ALPHA{}-rules here as well. The second and third kinds have no equivalents in standard tableau systems. We shall refer to the second kind as \SIGMA{}-rules, and to the branching rule as \BETA{}-rule -- after Smullyan's, even though his branching rules are different. \REF{fig:ExpansionRules} summarises the expansion rules to be employed by the system. A deduction bar says that if the formula(e) appearing above it occurs in the tree, then the formula(e) below it should be added to the tableau. The rules are easily interpreted according to the intuitions assigned above to signs, formulae and information tokens. A rule like \ALPHA{(i)}, for example, says that if A\BS B is not available for combination and x verifies such information, then this is because there is an A available at some token a, but the combination of a and x (notice that the order is relevant) does not produce B. \noindent \begin{figure}[!h] \centering\begin{tabular}{|l|l|l|l|l|l|l|} \hline \small{{\bf \ALPHA{}-rules}} & (i) & (ii) & (iii) & \multicolumn{2}{c}{\small{{\bf \BETA{}-rule}}} &\\ \hline \underline{(\ALPHA{1})} & \underline{F A\BS B : x} & \underline{F A/B : x} & \underline{\raisebox{-.1ex}[.05ex][.6ex]{ T A \BU B : x}} & \multicolumn{3}{c|}{} \\ (\ALPHA{2}) & T A : a$^{*}$ & T B : a$^{*}$ & T A : a$^{*}$ & \multicolumn{3}{c|}{\emth{\overline{(\BETA{1}) \hspace{0.2cm} T A : x \emth{\mid} (\BETA{2})\hspace{0.2cm} F A : x}}} \\ (\ALPHA{3}) & F B : a \CI x & F A : x \CI a & T B : x \opRD a & \multicolumn{3}{c|}{} \\ \hline \small{{\bf \SIGMA{}-rules}} & (i) & (ii) & (iii) & (iv) & (v) & (vi) \\ \hline (\SIGMA{1}) & T A\BS B : x & T A\BS B : x & T A/B : x & T A/B : x & F A \BU B : x & F A \BU B : x \\ \underline{(\SIGMA{2})} & \underline{T A : y} & \underline{F B : y \CI x} & \underline{F A: x \CI y} & \underline{T B : y} & \underline{T A : y} & \underline{T B : y} \\ (\SIGMA{3}) & TB : y \CI x & F A : y & F B : y & T A : x \CI y & F B : x \opRD y & F A : x \opLD y \\ \hline \end{tabular}\\ \emth{\ast}:{\em a new label {\em a} (not occurring previously in the derivation) must be introduced.} \caption{Tableau expansion rules}\label{fig:ExpansionRules} \end{figure} \noindent Given the expansion rules, the definition of the main data structure to be manipulated by the theorem proving (parsing) algorithm is straightforward: a derivation tree, \cT, is simply a binary tree built from a set of given formulae by applying the rules. The next step is to define the conditions for a tree to be regarded as complete. Completion along with inconsistency are the notions upon which the algorithm's termination depends. It can be readily seen on \REF{fig:ExpansionRules} that for a finite set of formulae, the number of times \ALPHA{} and \SIGMA{} rules can be applied increasing the number of SLFs (nodes) in \cT$\;$ is finite. Unbounded application of \BETA{}, however, might expand the tree indefinitely. In order to assure termination, applications of \BETA{} will be restricted to sub-formulae of formulae in \cT. These notions are formalised in \RED{def:Completion}. \DEF{(Tree Completion) Given \cT, a tree for a set of SLFs S, we say that a binary tree \cT * is a tableau for S if \cT * results from \cT$\;$ through the application of an expansion rule. A tableau \cT * is {\em linearly complete} if it satisfies the following conditions: (i) if \ALPHA{1} \EL \cT, then \ALPHA{2} and \ALPHA{3} \EL \cT; (ii) if \SIGMA{1} and \SIGMA{2} \EL \cT, then \SIGMA{3} \EL \cT. A tree \cT * is {\em complete} iff for every A \EL \cT * and every sub-formula A' of A, both F A' : x and T A' : x have been added to \cT* by an application of the \BETA{}-rule.}{def:Completion} Now, the first step towards building a counter-model for the denial of the formula to be proved is the search for a tree containing {\em potential} contradictions. Whether or not a potentially inconsistent tree is a counter-model for the formula will depend ultimately upon the constraints on the labelling algebra. This form of inconsistency is defined below. \DEF{(Branch and Tree Inconsistency) A branch is {\em inconsistent} iff for some type X both T X and F X, labelled by any information token, occur in the branch. A tree is inconsistent iff its branches are all inconsistent.}{def:TreeInconsistency} Given the definitions above, we are ready to define an algorithm for expanding linearly the derivation tree. For efficiency reasons non-branching rules will be exhaustively applied before we move on to employing \BETA{}-rules. \RED{def:LinearCompletion} presents the basic procedure for generating linear expansion for a branch.\footnote{The reader will notice that if we had allowed \SIGMA{2} formulae to search for \SIGMA{1} types for combination, in the same way that \SIGMA{1}s search for \SIGMA{2}s, then linear expansion would not terminate for some cases. Consider for example the infinite sequence of \SIGMA{} applications: T A / B: x, T B / A : y, T A : z, T B : y \CI z, T A : x \CI (y \CI z), T B : y \CI (x \CI (y \CI z)),... A strategy to allow unrestricted \SIGMA{}-application without running into non-terminating procedures, as well as other practical and computational issues is discussed in \cite{luz95c}.} The complete LKE algorithm, \RED{def:Expansion}, which uses the procedure below, will be presented after we have discussed tableau closure from the information frame perspective. \DEF{(Algorithm: Linear Completion) Given \cT, a LKE-tableau structure, we define the procedure: \vspace{-1.3ex}\textnormal{ \begin{tabbing} \mbox{{\sf Linear-Completion(\cT)}} \\ 99 \= \kill 1 \> \aDO \= \cT{} \MI \ALPHA{}-completion(\cT) \\ 2 \> \> formula \MI head[\cT] \\ 3 \> \aWHILE \= ( \NE completed(\cT) \= \aOR \= consistent(\cT) ) \\ 4 \> \> \aDO \= \aIF \= \SIGMA{1}-type(formula) \\ 5 \> \>\> \aTHEN \aDO \= formula\emth{_{aux}} \MI search(\cT,\SIGMA{2}) \= $\;\;$ \= \aCOMMENT{formula\emth{_{aux}} is a set of \SIGMA{2}-type slf's}\\ 6 \> \>\>\> \aIF \= formula\emth{_{aux}} \emth{\not = \emptyset} \> \aCOMMENT{\SIGMA{3}-set results from combining \SIGMA{1} to each \SIGMA{2} }\\ 7\> \>\>\>\> \aTHEN \aDO \= \SIGMA{3}-set \MI combine-labels(\SIGMA{1},formula\emth{_{aux}}) \\ 8\> \>\>\>\>\> \SIGMA{3}-expansion \MI \ALPHA{}-completion(\SIGMA{3}-set) \\ 9\> \>\>\>\>\> \cT \MI append(\cT,\SIGMA{3}-expansion) \\ 10\> \> \aDO formula \MI next[\cT] \\ 11\> \aRETURN \cT \end{tabbing} }}{def:LinearCompletion} We have seen above that the labels are means to propagate information about the formulae through the derivation tree. From a semantic viewpoint, the calculi addressed in this paper are obtained by varying the structure assigned to the set of formulae in the derivation\footnote{For instance, resource sensitive logics such as linear logic are frequently characterised in terms of multisets to keep track of the ``use'' of formulae throughout the derivation.}. Therefore, in order to verify whether a branch is closed for a calculus one has to verify whether the information frame satisfies the constraints which characterise the calculus. For instance, the standard Lambek calculus \calcL{} does not allow any sort of structural manipulation of formulae apart from associativity; \calcLP{} allows formulae to be permuted; \calcLPE{} allows permutations and expansion (i.e. if B can be proved from the sequent $\Delta$, A, $\Gamma$, then B can be proved from $\Delta$, A, A, $\Gamma$); \calcLPC{} allows permutation and contraction; etc. The definition below sets the algebraic counterparts of these properties. \DEF{An information frame is: (i) {\em associative} if x \CI (y \CI z) \PO (x \CI y) \CI z and (x \CI y) \CI z \PO x \CI (y \CI z); (ii) {\em commutative} if x \CI y \PO y \CI x; (iii) {\em contractive} if x \CI x \PO x; (iv) {\em expansive} if x \PO x \CI x; (v) {\em monotonic} if x \PO x \CI y, for all x, y, z \EL \cP.}{def:Constraints} Now, we say that a branch is {\em closed} with respect to the labelling algebra if it contains SLFs of the form T X : x and F X : y, where x \PO y. Likewise, a tree is closed if it contains only closed branches. Checking for label closure will depend on the calculus being used, and consists basically of reducing information token expressions to a {\em normal form}, via properties (\ref{Property1})--(\ref{Property4}), and then matching tokens and/or variables that might have been introduced by applications of the \BETA{}-rule according to the properties or combination of properties (\RED{def:Constraints}) that characterise the calculus considered. It should be noticed that, in addition to the basic algorithm, heuristics might be employed to account for specific linguistic aspects. Some examples: (a) it could be assumed that all the bracketing for the strings is to the right thus favouring an incremental approach; (b) type reuse could be blocked at the level of the formulae, reducing the the computational cost of searches for label closure, since most of the calculi in the family covered by the system are resource sensitive; (c) priority could be given to juxtaposed strings for \SIGMA{}-rule application, etc. \RED{def:Expansion} gives the general procedure for tableau expansion, abstracted from the heuristics mentioned above. \DEF{(Algorithm: LKE-completion) The complete tableau expansion for a LKE-tree \cT is given by the following procedure: \textnormal{ \begin{pseudocode}{expansion(\cT)} 1 \> \aDO closure-flag \MI no \\ 2 \> \aWHILE \= \NE ( completed(\cT) \= \aOR \= closed-\cT = yes) \\ 3 \>\> \aDO \= \cT \MI linear-completion(\cT) \\ 4 \> \>\> \aIF \= \NE consistent(\cT) \aAND label-closure(\cT) \\ 5 \> \>\>\> \aTHEN \aDO \= closure-flag \MI yes \\ 6 \> \>\>\> \aELSE \aDO \= subformula \MI select-subformula(\cT) \\ 7 \> \>\>\>\> subformula\emth{_{T}} \MI assign-label-T(subformula)\\ 8 \> \>\>\>\> subformula\emth{_{F}} \MI assign-label-F(subformula) \\ 9 \> \>\>\>\> \cT\emth{_{1}} \MI append(\cT,\{subformula\emth{_{T}}\}) \\ 10\> \>\>\>\> \cT\emth{_{2}} \MI append(\cT,\{subformula\emth{_{F}}\}) \\ 11\> \>\>\>\> \aIF \= ( expansion(\cT\emth{_{1}}) = yes \aAND expansion(\cT\emth{_{2}}) = yes ) \\ 12\> \>\>\>\>\> \aTHEN \aDO closure-flag \MI yes \\ 13\> \aRETURN closure-flag \end{pseudocode}}}{def:Expansion} As it is, the algorithm defined above constitutes a semi-decision procedure. This is due to the fact that even though the search space for signed formulae is finite, the search space for the labels is infinite. The labels introduced via \BETA{}-rules are in fact universally quantified variables which must be instantiated during the label unification step. This represents no problem if we are dealing with theorems, i.e. trees which actually close. However, for completed trees with an open branch, the task might not terminate. In order to overcome this problem and bind the unification procedure we restrict label (variable) substitutions to the set of tokens occurring in the derivation --- similarly to the way parameter instantiation is dealt with by liberalized quantification rules for first-order logic tableaux. In practice, the strategy adopted to reduce label complexity also employs the following refinements: (i) the tableau is linearly expanded keeping track of the choices made when \SIGMA{}-rules are applied (the options are kept in a stack); (ii) once this first step is finished, if the tableau is still open, then backtrack is performed until either the choices left over are exhausted or closure is achieved; (iii) only then is the \BETA{}-rule applied. This explains the role played by the heuristics mentioned above.\footnote{Furthermore, as we will show in section \ref{sec:Comparison}, if associativity is allowed at the syntactic level then it is possible to eliminate the branching rule for the class of calculi discussed here.} We are now able to establish some results regarding the reduction laws mentioned in section \ref{sec:Family}. \begin{proposition1}[Reduction Laws] Let X, Y and Z be types, and \cL an information frame. The properties R1--R5 hold: \end{proposition1} \begin{proof} \label{prf:ReductionLaws} The proofs are obtained by straightforward application of \RED{def:LinearCompletion} and \RED{def:Expansion}. Below we illustrate the method by proving (R1) and (R2): \begin{itemize} \item[(R1)] To prove right application we start by assuming that it is verified by the identity token \ID{}. From this we have: 1- T X/Y \BU Y : m, 2- F X : 1 \CI m = m. Then, we apply \ALPHA{(iii)} to 1 obtaining 3- T X/Y : n and 4- T Y : m \opRD n. The next step is to combine 3 and 4 via \SIGMA{(iv)} getting 5- T X : n \CI (m \opRD n). Now we have a potential closure caused by 5 and 2. If we apply property (\ref{Property1}) to the label for 5 we find that n \CI (m \opRD n) \PO m, which satisfies the closure condition thus closing the tableau. \item[(R2)] Let's prove left composition. As we did above, we start with: 1- T Z\BS Y \BU Y\BS X : m and 2- F Z\BS X : 1 \CI m. Applying \ALPHA{(iii)} to 1 we get: 3- T Z\BS Y : a and 4- T Y\BS X : m \opRD a. Now, we may apply \ALPHA{(i)} to 2 and get: 5- T Z : b and 6- F X : b \CI m. Then, combining 3 and 5 via \SIGMA{(i)}: 7- T Y : b \CI a. And finally 4 and 7 through the same rule: 8- T X : (b \CI a) \CI (m \opRD a). The closure condition for 8 and 6 is achieved as follows: \vspace{-1ex} \[ \begin{array}{llll} (b \CI a) \CI (m \opRD a) & \PO b \CI (a \CI (m \opRD a)) & \textit{by associativity} \nonumber \\ & \PO b \CI m & \textit{by (\ref{Property1}) and \CI being order-preserving} &\nonumber \end{array} \] \end{itemize} \vspace{-4ex}\hfill\QED \end{proof} Even though \calcL{} does not enjoy finite axiomatizability, the results above suggest that the calculus finds a natural characterization in LKE for associative information frames. In particular, the Division Rule ({\em R6}) can be regarded as \calcL{}'s characteristic theorem, since it is not derivable in weaker calculi such as \calcAB{}, \calcNL{}, and \calcF{}. If we do not allow associative frames, we get \calcNL{}. Stronger calculi such as \calcLP{}, \calcLPE{}, \calcLPC{} and \calcLPCE{} \cite{moortgat93} can be obtained for the same general framework by assigning further properties to \CI in the labelling algebra. Frames exhibiting combinations of monotonicity, expansivity, commutativity and contraction allow us to characterise these substructural calculi. Algebras that are both associative and commutative describe \calcLP{}. Adding expansivity (weakening) to \calcLP{} results in \calcLPE{}. Associativity, commutativity and contraction describe \calcLPC{} frames. \calcLPCE{} is obtained by combining the properties of \calcLPC{} and \calcLPE{} algebras. We end this section with a simple example requiring associativity: show, in \calcL{}, that an NP ({\sf John}), combined with a type (NP\BS S)/NP ({\sf likes}) yields S/NP, i.e a type which combined with a NP will result in a sentence (Proof \ref{prf:JLBBHM}). \noindent\begin{proof}Let's assume the following type--string correspondence: NP for {\sf John}, (NP\BS S)/NP for {\sf likes}. The expression we want to find a counter-model for is: 1- F NP \BU (NP\BS S)/NP \ENI{\! L} S/NP. Therefore, the following has to be proved: 2- T NP \BU (NP\BS S)/NP : m and 3- F S/NP : m. We proceed by breaking 2 and 3 down via \ALPHA{(iii)}, obtaining: 4- T NP : a, 5- T (NP\BS S)/NP : m\opRD a, 6- T NP : b, and 7- F S : (m \CI b). \\ \noindent Now we start applying \SIGMA{}-rules (annotated on the right-hand side of each line):\\ \noindent\begin{tabular}{llll} 8- T & NP\BS S & : (m \opRD a) \CI b & 5,6 \SIGMA{(i)} \\ 9- T & S & : a \CI ((m \opRD a) \CI b) & 4,9 \SIGMA{(i)} \\ \end{tabular} \\ We have derived a potential inconsistency between 7 and 9. Turning our attention to the information tokens, we verify closure for \calcL{} as follows: \begin{center} \begin{tabular}{lllll} a \CI ((m \opRD a) \CI b) & \PO & (a \CI (m \opRD a)) \CI b & by associativity \\ & \PO & m \CI a & by property (\ref{Property1}) \end{tabular}\end{center} \vspace{-3ex} \hfill \QED \label{prf:JLBBHM} \end{proof} \subsection{Comparison with Existing Approaches} \label{sec:Comparison} Early implementations of CG parsing relied on cut-free Gentzen sequents implemented via backward chaining mechanisms \cite{moortgat88}. Apart from the fact that it lacks generality, since implementing more powerful calculi would involve modifying the code in order to accommodate new structural rules, this approach presents several sources of inefficiency. The main ones are: the generate-and-test strategy employed to cope with associativity, the non-determinism in the branching rules and in rule application itself. The impact of the latter form of non-determinism over efficiency can be reduced by testing branches for {\em count invariance} prior to their expansion and by performing sequent proof normalisation. However, non-determinism due to splitting in the proof structure still remains. As we move on to stronger logics and incorporate structural modalities such problems tend to get even harder. An improved attempt to deal uniformly with multiple calculi is presented in \cite{moortgat92}. In that paper, the theorem prover employed is based on proof nets, and the characterisation of different calculi is taken care of by labelling the formulae. For substructural calculi stronger than \calcL{}, much of the complexity (perhaps too much) is shifted to the label unification procedures. A strategy for improving such procedures by compiling labels into higher-order logic programming clauses is presented in \cite{morrill95} for \calcNL{} and \calcL{}. However, a comprehensive solution to the problem of binding label unification, a problem which arises as we move from sequents to labelled proof nets, has not been presented yet. Moreover, as discussed in \cite{leslie90}, if we consider that the system is to be used as a parser, as a tool for linguistic study, the proof net style of derivation does not provide the clearest or most intuitive display of the proofs.\footnote{Proof nets and sequent normalisation have also been employed to get around spurious ambiguity (i.e. multiple proof for the same sentence, with the same semantics). Our approach does not exhibit this problem.} In our approach, the burden of parsing is not so concentrated in label unification but is more evenly divided between the theorem prover and the algebraic checker. This is mainly due to the fact that the system allows for a controlled degree of non-determinism, present in the \SIGMA{}-rules, which enables us to reduce the introduction of variables in the labelling expressions to a minimum. We believe this represents an improvement on previous attempts. Besides this, controlling composition via bounded backtrack opens the possibility of implementing heuristics reflecting linguistic and contextual knowledge. In fact, we verify that, under the appropriate application of rules, we are able to eliminate the \BETA{}-rule for a class of theorems. \begin{proposition1}[Elimination Theorem] All closed LKE-trees derivable by the application of the set of rules \cR = \SE{\ALPHA{(i)},...,\ALPHA{(iii)},\SIGMA{(i)},...,\SIGMA{(vi)},\BETA{}} can be also derived from \cR $-$ \{\BETA{}\} + \{ assoc \}. \end{proposition1} The proof of this proposition can be done by defining an {\em abstract Gentzen relation}, proving a substitution lemma with respect to the labelling algebra (as in \cite{dagostino94}), and showing that our consequence relation is closed under the relevant Gentzen conditions even if no \BETA{} rule is employed. The proof appeals to the fact that no formula signed by F can occur in the sequents on the left-hand side of the entailment relation, since the calculi presented here do not have negation.\footnote{Of course, without the \BETA{} rule not all open trees generated will constitute downward saturated sets, since they might contain formulae which are not completely analysed.} We believe that this result shows that, even though LKE label unification might be computationally expensive for substructural logics in general, the system seems to be well suited for categorial logics. We refer the reader to \cite{luz95c} for a more comprehensive discussion of these issues. \section{Conclusions and Further Work} \label{sec:Conclusion} We have described a framework for the study of categorial logics with different degrees of expressivity on a uniform basis, providing a tool for testing the adequacy of different CGs to a variety of linguistic phenomena. From a practical point of view, we have investigated the effectiveness and generality issues of a parsing strategy for CG opening an avenue for future developments. Moreover, we have pointed out some strategies for improving on efficiency and for dealing with more expressive languages, including structural modalities. The architecture proposed seems promising. Its flexibility with respect to the variety of logics it deals with, and its modularity suggest some natural extensions to the present work. Among them: implementing a semantic module based on Curry-Howard correspondence between type deduction and $\lambda$-terms, adding local control of structural transformations (structural modalities) to the language, increasing expressivity in the information frames for covering calculi weaker than \calcL{} (e.g. Dependency Categorial Grammar \cite{pickering93}), exploiting the derivational structure encoded in the labels to define heuristics for models of human attachment preferences etc. Problems for further investigation might include: the treatment of polymorphic types (by incorporating rules for dealing with quantification analogous to Smullyan's \gD and \gG rules \cite{fitting90a} \cite{smullyan68}), and complexity issues regarding how the general architecture proposed here would behave under more standard theorem proving methods.
\chapter{Introduction} Despite their central role in string compactifications, chiral $(0,q)$ supersymmetric sigma models have received less attention in comparison to the non chiral models due to difficulties involved with their construction and quantization. Recently there has been considerable interest in the use of massive linear sigma models to construct string vacua as the infrared conformal fixed point of the renormalization group flow. This method allows one to construct a large class of string vacua including those with chiral supersymmetry. In particular an interesting paper by Witten [\Wittena] discusses a class of massive linear sigma models possessing on-shell (0,4) supersymmetry which flow in the infrared to conformally invariant sigma models describing ADHM instantons [\ADHM,\WW]. Previous work on the ADHM sigma model has focused primarily on classical aspects of the $(0,4)$ supersymmetry multiplet used and in particular the construction of off-shell superfield formalisms [\GR,\GS,\GSt]. Here we will study the models quantum properties and it's rich interplay between geometry and field theory in detail. The general $(p,q)$ supersymmetric massive sigma model has been constructed before [\HPT,\PT] and it's quantization is discussed to two loop order in [\NDL]. We will show here that the ADHM sigma model is ultraviolet finite to all orders of perturbation theory and integrate out the massive fields to obtain the low energy effective theory. Due to anomalies this theory has interesting non trivial properties and we obtain the quantum corrections to order $\alpha'^2$ by requiring that the anomalies are appropriately canceled. We conclude by making some comments about the case where the instanton size vanishes. \chapter{The ADHM Sigma Model} In [\Wittena] Witten constructs an on-shell (0,4) supersymmetric linear sigma model which parallels the ADHM construction of instantons [\ADHM]. The model consists of $4k$ bosons $X^{AY}$, $A=1,2,\ Y=1,2...,2k$ with right handed superpartners $\psi^{A'Y}_-$, $A'=1,2$. There is also a similar multiplet of fields $\phi^{A'Y'},\ \chi_-^{AY'}$ $Y'=1,2...,2k'$. In addition there are $n$ left handed fermions $\lambda^a_+$, $a=1,2...,n$. The $A,B...$ and $A',B'...$ indices are raised (lowered) by the two by two antisymmetric tensor $\epsilon^{AB}$ ($\epsilon_{AB}$), $\epsilon^{A'B'}$ ($\epsilon_{A'B'}$). The $Y,Z...$ and $Y',Z'...$ indices are raised (lowered) by the invariant tensor of $Sp(k)$, $Sp(k')$ respectively which are also denoted by $\epsilon^{YZ}$ ($\epsilon_{YZ}$), $\epsilon^{Y'Z'}$ ($\epsilon_{Y'Z'}$). The interactions are provided for by a tensor $C^a_{AA'}(X,\phi)$ in a similar manner to the construction of the general models [\HPT,\PT]. The action for the theory is given by $$\eqalign{ S = \int\! d^2x & \left\{ \epsilon_{AB}\epsilon_{YZ} {\partial}_= X^{AY} {\partial}_{\ne} X^{BZ} +i\epsilon_{A'B'}\epsilon_{YZ} \psi_-^{A'Y} {\partial}_{\ne} \psi_-^{B'Z} \right. \cr & \left. +\epsilon_{A'B'}\epsilon_{Y'Z'}{\partial}_= \phi^{A'Y'} {\partial}_{\ne} \phi^{B'Z'} +i\epsilon_{AB}\epsilon_{Y'Z'}\chi_-^{AY'} {\partial}_{\ne} \chi_-^{BZ'} \right. \cr & \left. +i\lambda_+^a {\partial}_= \lambda^a_+ -{im\over2}\lambda_+^a \left(\epsilon^{BD}{\partial C^a_{BB'}\over \partial X^{DY}}\psi_-^{B'Y} +\epsilon^{B'D'}{\partial C^a_{BB'}\over \partial \phi^{D'Y'}}\chi_-^{BY'}\right) \right. \cr & \left. -{m^2\over8}\epsilon^{AB}\epsilon^{A'B'}C^a_{AA'}C^a_{BB'} \right\}\ , \cr} \eqn\action $$ where $$ {\partial}_{\ne} = {1\over{\sqrt{2}}}(\partial_0 + \partial_1 )\ \ \ \ \ \ \ \ \ \ {\partial}_= = {1\over{\sqrt{2}}}(\partial_0 - \partial_1 )\ , $$ and $m$ is an arbitrary mass parameter. Note the twisted form of the Yukawa interactions in \action\ in comparison to the models of [\HPT,\PT]. The free field theory ($m$ =0) possesses an $SU(2)\times Sp(k)\times SU(2) \times Sp(k')$ rigid symmetry acting on the $AB, YZ,A'B', Y'Z'$ indices respectively which is generally broken by the potential terms. Provided that $C^a_{AA'}$ takes the simple form ($M^a_{AA'}$,$N^a_{A'Y}$,$D^a_{AY'}$ and $E^a_{YY'}$ are constant tensors) $$ C^a_{AA'} = M^a_{AA'} + \epsilon_{AB}N^a_{A'Y} X^{BY} + \epsilon_{A'B'}D^a_{AY'}\phi^{B'Y'} + \epsilon_{AB}\epsilon_{A'B'}E^a_{YY'} X^{BY} \phi^{B'Y'} \ , \eqn\onesusy $$ subject to the constraint $$ C^a_{AA'}C^a_{BB'}+C^a_{BA'}C^a_{AB'} = 0 \ , \eqn\foursusy $$ then the action \action\ has the on-shell (0,4) supersymmetry $$\eqalign{ \delta X^{AY} &= i\epsilon_{A'B'}\eta^{AA'}_+\psi_-^{B'Y}\cr \delta \psi^{A'Y}_- &= \epsilon_{AB}\eta^{AA'}_+{\partial}_= X^{BY}\cr \delta \phi^{A'Y'} &= i\epsilon_{AB}\eta^{AA'}_+\chi_-^{BY'}\cr \delta \chi^{AY'}_- &= \epsilon_{A'B'}\eta^{AA'}_+{\partial}_= \phi^{B'Y'} \cr \delta \lambda^{a}_+ &= \eta^{AA'}_+C^a_{AA'} \ ,\cr} \eqn\susy $$ where $\eta^{AA'}_+$ is an infinitesimal anticommuting spinor parameter. As is discussed by Witten [\Wittena], the above construction of models with (0,4) supersymmetry can be interpreted as a string theory analogue of the ADHM construction of instantons with instanton number $k'$ in a target space dimension of $4k$. The general form of massive $(p,q)$ supersymmetric sigma models has been discussed in terms of $(0,1)$ superfields in [\HPT,\PT] and we now provide such a formulation of the ADHM model. To this end we introduce a tensor $I^A_{\ A'}$ satisfying $$ \epsilon_{AB}I^A_{\ A'}I^B_{\ B'} = \epsilon_{A'B'} \eqn\Idef $$ which can be interpreted as a complex structure in the sense that $I^{AB'}I_{AC'}=-\delta^{B'}_{C'}$, $I^{BA'}I_{CA'}=-\delta^{B}_{C}$. The 'twisted' superfields are $$\eqalign{ {\cal X} ^{AY} &= X^{AY} + \theta^-I^A_{\ A'}\psi_-^{A'Y} \cr \Phi^{A'Y'} &= \phi^{A'Y'} + \theta^-I^{\ A'}_{A}\chi_-^{AY'}\ \cr \Lambda^a_+ &= \lambda^a_+ + \theta^-F^a \ ,\cr} \eqn\superfields $$ where $\theta^-$ is the anticommuting spinorial (0,1) superspace coordinate with the associated superspace covariant derivative $$ D_- = {\partial\over\partial \theta^-} + i\theta^-{\partial}_= $$ and $F^a$ is an auxiliary field. After removing $F^a$ by it's equation of motion and using the constraint \foursusy, the action \action\ can be seen to have the superspace form $$\eqalign{ S_{effective} = -i\int\! d^2xd \theta^- & \left\{ \epsilon_{AB}\epsilon_{YZ} D_-{\cal X}^{AY}{\partial}_{\ne} {\cal X}^{BZ} +\epsilon_{A'B'}\epsilon_{Y'Z'} D_-\Phi^{A'Y'}{\partial}_{\ne} \Phi^{B'Z'} \right. \cr &\left. - i \delta_{ab}\Lambda^a_+ D_- \Lambda^b_+ - mC_a\Lambda^a_+ \right\} \ , \cr} \eqn\suspaction $$ where $C^a = I^{AA'}C^a_{AA'}$. Since the vector field $C^a$ is harmonic and the target space is ${\bf R}^{4(k+k')}$, the model satisfies the requirements of $(0,4)$ supersymmetry found in [\PT]. The inclusion of the auxiliary field allows one to close a $(0,1)$ part of the supersymmetry algebra off-shell. As with the component field formulation \action\ the full $(0,4)$ supersymmetry is only on-shell. A manifestly off-shell form requires harmonic superfields with an infinite number of auxiliary fields [\GS]. Lastly we outline the $k=k'=1,\ n=8$ case (ie. a single instanton in ${\bf R}^4$) analyzed by Witten which will be of primary interest here. The right handed fermions are taken to be $\lambda^a_+ = (\lambda^{AY'}_+,\lambda^{YY'}_+)$ and the tensor $C^a_{AA'}$ takes the form $$\eqalign{ C^{YY'}_{BB'} &= \epsilon_{BC}\epsilon_{B'C'}X^{CY}\phi^{C'Y'} \cr C^{AY'}_{BB'} &= {\rho\over\sqrt2}\epsilon_{B'C'}\delta^A_B\phi^{C'Y'}\ , \cr} $$ where $\rho$ is an arbitrary constant to be interpreted as the instanton size. The bosonic potential for this theory is easily worked out as $$ V = {m^2\over8}(\rho^2 + X^2)\phi^2 \ , \eqn\V $$ where $X^2 = \epsilon_{AB}\epsilon_{YZ}X^{AY}X^{BZ}$ and similarly for $\phi^2$. Thus, for $\rho \ne 0$, the vacuum states of the theory are defined by $\phi^{A'Y'}=0$, and parameterize ${\bf R}^4$. The $X^{AY}$ and $\psi_-^{A'Y}$ are massless fields while $\phi^{A'Y'}$ and $\chi_-^{AY'}$ are massive. This yields exactly 4 of the $\lambda^a_+$ massive and 4 massless. \chapter{Quantization} \subsection{Renormalization} It is not hard to see that the model described above is superrenormalizable in two dimensions as the interaction vertices do not carry any momentum factors and have at most three legs. In fact a little inspection reveals that the only possible divergences of the theory are the one loop graphs contributing to the potential. This can also been seen from the following simple superspace argument, valid for any $(0,1)$ supersymmetric linear massive sigma model. The superspace measure $d^2xd\theta^-$ has mass dimension $-3/2$ while all vertices contribute a factor of $m$ to the effective action. Thus by power counting, only graphs with a single vertex can yield divergent contributions to the effective action. Of these, only the one loop (tadpole) graphs are relevant, all the higher loop divergences are removed by the renormalization procedure. Although the potential provides masses for some of the fields there will in general be massless fields which may cause infrared divergences. We must therefore add an infrared regulator in the form of a mass $M$ to the propagator and treat any mass terms in \action\ as interactions, taking the limit $M\rightarrow 0$ in the final expressions. Using dimensional regularization in $D=2+\epsilon$ dimensions and the background field method [\Back] the bosonic graphs are readily calculated to be $$\eqalign{ \Gamma_{Div}({\rm bosons})=-&{m^2\over4}\Delta(0) \left[\epsilon^{AB}\epsilon^{CD}\epsilon^{C'D'}\epsilon^{YZ} {\partial C^a_{CC'}\over\partial X^{AY}} {\partial C^a_{DD'}\over\partial X^{BZ}} \right.\cr & \left. \ \ \ \ \ \ +\epsilon^{A'B'}\epsilon^{C'D'}\epsilon^{CD}\epsilon^{Y'Z'} {\partial C^a_{CC'}\over\partial \phi^{A'Y'}} {\partial C^a_{DD'}\over\partial \phi^{B'Z'}} \right] \ , \cr} \eqn\bosons $$ while the fermionic graphs are $$\eqalign{ \Gamma_{Div}({\rm fermions}) =&{m^2\over4}\Delta(0) \left[{1\over2}\epsilon^{AC}\epsilon^{BD}\epsilon^{C'D'}\epsilon^{YZ} {\partial C^a_{CC'}\over\partial X^{AY}} {\partial C^a_{DD'}\over\partial X^{BZ}} \right. \cr & \left. \ \ \ \ \ \ +{1\over2}\epsilon^{A'C'}\epsilon^{B'D'}\epsilon^{CD}\epsilon^{Y'Z'} {\partial C^a_{CC'}\over\partial \phi^{A'Y'}} {\partial C^a_{DD'}\over\partial \phi^{B'Z'}} \right]\ , \cr} \eqn\fermions $$ where all the tensor expressions are evaluated at the background fields and the bosonic propagator at zero momentum and renormalization scale $\mu$ is $$ \Delta (0) = -{1\over 2\pi\epsilon}-{\rm ln}\left({M^2\over\mu^2}\right) +{\rm finite} \ . $$ One can see that the epsilon tensor terms in \bosons\ and \fermions\ are different as a result of the twisted form of the Yukawa interactions. It is not immediately obvious then that the bosonic and fermionic divergences cancel. However by substituting in \onesusy\ it is not much trouble to see that they do and hence $\Gamma_{Div}=0$. Thus Witten's ADHM model is ultraviolet finite to all orders of perturbation theory. Therefore there is no renormalization group flow in these models. This result may be expected, but is not guaranteed, by supersymmetry as there is a general argument for finiteness only for off-shell $(0,4)$ sigma models, with some modifications required due to anomalies [\HPb]. \subsection{Integrating the Massive Modes} In this section we will integrate out the massive modes. We shall postpone the problem of anomalies in chiral supersymmetric models until the next section. We assume for simplicity here that $M^a_{AA'}=N^a_{AY}=0$, $D^a_{A'Y'} \ne 0$ so that the $X^{AY}$ and $\psi^{A'Y}_-$ fields are massless, the $\phi^{A'Y'}$ and $\chi_-^{AY'}$ fields massive and the vacuum is at $\phi^{A'Y'}=0$. The theory is then only quadratic in the massive fields and integrating over them is therefore exact at the one loop level. This assumption also ensures that the interacting theory breaks the $SU(2) \times Sp(k) \times SU(2) \times Sp(k')$ symmetry of the free theory down to $SU(2) \times Sp(k')$ [\Wittena]. We may therefore write $$\eqalign{ C^a_{AA'} &=\epsilon_{A'B'}D^a_{AY'}\phi^{B'Y'} + \epsilon_{AB}\epsilon_{A'B'}E^a_{YY'} X^{BY}\phi^{B'Y'} \cr &\equiv \epsilon_{A'B'}B^a_{AY'}(X)\ \phi^{B'Y'} \ . \cr} $$ At this point it is necessary to split up the left handed fermions into there massive and massless parts. If we introduce the zero modes $v^a_i(X)$, $i=1,2,...,n-4k'$ of the fermion mass matrix, defined such that $$ v^a_iB^a_{AY'} = 0, \ \ \ \ v^a_iv^a_j = \delta_{ij} $$ and a similar set of massive modes $u^a_I(X)$, $I=1,2,...,4k'$ satisfying $$ u^a_Iu^a_J = \delta_{IJ}, \ \ \ \ u^a_Iv^a_j =0 $$ then we may separate the $\lambda^a_+$ as $$ \lambda^a_+ = v^a_i\zeta^i_+ + u^a_I\zeta^I_+ \ . \eqn\lamodes $$ so that the $\zeta^i_+$ are massless and the $\zeta^I_+$ massive. We now rewrite the action \action\ in terms of the massless and massive fields $$ S = S_{massless} + S_{massive} \eqn\S $$ where $S_{massless}$ is the part of \action\ which only contains the massless fields. Explicitly $$ \eqalign{ S_{massless} = \int\! d^2x & \left\{ \epsilon_{AB} \epsilon_{YZ} {\partial}_= X^{AY} {\partial}_{\ne} X^{BZ} + i \epsilon_{A'B'} \epsilon_{YZ} \psi^{A'Y}_- {\partial}_{\ne} \psi^{B'Z}_- \right. \cr &\left. +i\zeta^i_+ (\delta_{ij}{\partial}_=\zeta^j_+ + A_{ijAY}{\partial}_= X^{AY}\zeta^j_+) \right\} \ ,\cr} \eqn\Smassless $$ where $$ A_{ijAY}=v^a_i{\partial v^a_j\over\partial X^{AY}} \eqn\Adef $$ is the induced $SO(n-4k')$ connection and $$\eqalign{ S_{massive} = \int\! & d^2x \left\{ \epsilon_{A'B'} \epsilon_{Y'Z'} {\partial}_=\phi^{A'Y'}{\partial}_{\ne}\phi^{B'Z'} + i\epsilon_{AB} \epsilon_{Y'Z'} \chi^{AY'}_- {\partial}_{\ne} \chi^{BZ'}_- \right. \cr & \left. +i\delta_{IJ}\zeta^I_+{\partial}_= \zeta^J_+ + iA_{IJAY} {\partial}_= X^{AY} \zeta^I_+ \zeta^J_+ +2iA_{iJAY} {\partial}_= X^{AY} \zeta^i_+\zeta^J_+ \right. \cr & \left. -im\epsilon_{B'C'}v^a_iE^a_{YY'}\zeta^i_+\phi^{C'Y'}\psi^{B'Y}_- -im\epsilon_{B'C'}u^a_IE^a_{YY'}\zeta^I_+\phi^{C'Y'}\psi^{B'Y}_- \right. \cr & \left. -imu^a_{I}B^a_{BY'}\zeta^I_+\chi_-^{BY'} -{m^2\over8}\epsilon^{AB}\epsilon_{C'D'} B^a_{AY'}B^a_{BZ'}\phi^{C'Y'}\phi^{D'Z'} \right\} \ ,\cr} \eqn\Smassive $$ where $A_{IJAY}=u^a_I{\partial u^a_J}/ {\partial X^{AY}}$ and $A_{iJAY}=v^a_i{\partial u^a_J}/ {\partial X^{AY}}$. The classical low energy effective action is simply obtained by considering the most general action possible which is compatible with all of the symmetries of the theory. To calculate the effective action quantum mechanically we will integrate over the massive fields and discard any higher derivative terms. The presence of higher derivative terms in the effective action, which are suppressed by powers $p/m$ where $p$ is the low energy momentum scale, would ruin the renormalizability and prevent a simple geometrical sigma model interpretation of the effective theory. First we notice that because of the nontrivial definition of the massless left handed fermions \lamodes , $S_{massless}$ is not (0,4) supersymmetric by itself as it is missing a four fermion interaction term. The problem is rectified by noting that there is a tree graph, with a single internal $\phi^{A'Y'}$ field propagating, which contributes to the low energy effective action. In order to avoid the singular behaviour of the propagator at zero momentum, when calculating this graph it is helpful to use the massive propagator for $\phi^{A'Y'}$, obtained from the last term in \Smassive . At this point it is useful to write, using \foursusy\ , $$ B^a_{AY'}B^a_{BZ'} = \epsilon_{AB}\epsilon_{Y'Z'}\Omega $$ where $$ \Omega(X)={1\over4k'}\epsilon^{AB}\epsilon^{Y'Z'}B^a_{AY'}B^a_{BZ'} \ . \eqn\Om $$ The last term in \Smassive\ becomes $$ -{m^2\over4}\Omega(X)\ \phi^2 \ , $$ and hence can be interpreted as the $X^{AY}$ dependent mass term for $\phi^{A'Y'}$. The tree graph can then be seen to contribute the four fermion term $$ -{1\over2}\zeta^i_+ \zeta^j_+F^{ij}_{A'YB'Z}\psi^{A'Y}_- \psi^{B'Z}_- $$ where $$ F^{ij}_{A'YB'Z} = 2 \epsilon_{A'B'} \epsilon^{Y'Z'}\Omega^{-1} v^a_i E^a_{(Y|Y'} v^b_j E^b_{|Z)Z'} \ , \eqn\Fdef $$ which we will later relate to the field strength tensor. We may now discard all vertices with only one massive field in \Smassive\ and examine the one loop contributions to the effective action. Inspection of the quadratic terms in $S_{massive}$ shows there are no contributions to the gauge connection in \Smassless. Furthermore, inspection shows that of all the other possible contributions only those corresponding to the effective potential do not involve higher order derivatives of the massless fields. A check on this is to note that any terms which are second order in the derivatives are logarithmically divergent, and by finiteness of the model, must therefore vanish. To calculate the effective potential we simply set ${\partial}_= X^{AY} = {\partial}_{\ne} X^{AY}= \psi^{A'Y}_-=0$. Thus only the last two terms in \Smassive\ need be considered (we no longer use a massive propagator for $\phi^{A'Y'}$). The effective potential then receives the standard bosonic and fermionic contributions (in Euclidean momentum space) $$ V_{eff}({\rm bosons}) = {\alpha'\over4\pi}\sum_{n=1}^{\infty}{1\over n} \int\! d^2 p\ {\rm Tr}\left[{\epsilon_{C'D'} \epsilon^{AB}B^a_{AY'}B^a_{BZ'}\over 4p^2/m^2}\right]^n \eqn\Veffbose $$ and $$ V_{eff}({\rm fermions}) = -{\alpha'\over4\pi} \sum_{n=1}^{\infty}{1\over n} \int\! d^2 p\ {\rm Tr}\left[{ u^a_Iu^{bI}B^a_{CY'}B^b_{DZ'}\over 2 p^2/m^2} \right]^n \ . \eqn\Vefffermi $$ Now the definition \Om\ yields the following expressions $$ \epsilon_{C'D'}\epsilon^{AB}B^a_{AY'}B^a_{BZ'} = 2\epsilon_{C'D'}\epsilon_{Y'Z'}\Omega $$ and $$ u^a_Iu^{bI}B^a_{CY'}B^b_{DZ'} = \epsilon_{CD}\epsilon_{Y'Z'}\Omega \ . $$ Therefore \Veffbose\ is completely canceled by \Vefffermi\ and there is no contribution to the effective potential. {}From the above analysis we conclude that the effective quantum action of the massless fields is $$\eqalign{ S_{effective} & = \int\! d^2x \left\{ \epsilon_{AB} \epsilon_{YZ} {\partial}_= X^{AY} {\partial}_{\ne} X^{BZ} + i\epsilon_{A'B'} \epsilon_{YZ} \psi^{A'Y}_- {\partial}_{\ne} \psi^{B'Z}_- \right. \cr &\left. \ \ \ +i\zeta^i_+ (\delta_{ij}{\partial}_= \zeta^j_+ + A_{ijAY}{\partial}_= X^{AY}\zeta^j_+) - {1\over2}\zeta^i_+ \zeta^j_+F^{ij}_{A'YB'Z}\psi^{A'Y}_- \psi^{B'Z}_- \right\} \ .\cr} \eqn\effective $$ This is simply the action of the general (0,4) supersymmetric nonlinear sigma model [\HPT,\PT], although the right handed superpartners of $X^{AY}$ are 'twisted'. As with the original theory \action , the low energy effective theory \effective\ admits a $(0,1)$ superfield form. Introducing the superfield $\Lambda^i_+ =\zeta^i_+ + \theta^-F^i$ then allows us (after removing $F^i$ by it's equation of motion) to express \effective\ as $$\eqalign{ S_{effective} = -i\int\! d^2xd & \theta^- \left\{ \epsilon_{AB}\epsilon_{YZ} D_-{\cal X}^{AY}{\partial}_{\ne} {\cal X}^{BZ} \right. \cr & \left.\ \ \ \ \ - i\Lambda^i_+ (\delta_{ij} D_- \Lambda^j_+ + A_{ijAY} D_-{\cal X}^{AY}\Lambda^j_+) \right\} \ , \cr} \eqn\superaction $$ provided that $F^{ij}_{A'YB'Z}$ satisfies $$ F^{ij}_{A'YB'Z} = I^A_{\ A'}I^B_{\ B'}F^{ij}_{AYBZ} \eqn\constraint $$ where $F^{ij}_{AYBZ}$ is the curvature of the connection \Adef , $$ F^{ij}_{AYBZ} = \partial_{AY} A_{ijBZ} - \partial_{BZ} A_{ijAY} + A_{ikAY}A_{kjBZ}- A_{ikBZ} A_{kjAY}\ . $$ This is just the familiar constraint on (0,4) models that the field strength be compatible with the complex structure [\HPT,\PT]. Furthermore it is not hard to check that $S_{effective}$ does indeed possess the full on-shell (0,4) supersymmetry (the superspace formulation \superaction\ ensures only off-shell (0,1) supersymmetry) precisely when \constraint\ is satisfied. For the $k=k'=1,\ n=8$ model above it is straightforward to determine the non zero components of $v^a_i$ and $u^a_I$ as $$ \eqalign{ v^{YY'}_{ZZ'}& = {\rho\over\sqrt{\rho^2+X^2}}\delta^Y_Z \delta^{Y'}_{Z'}\ \ \ \ \ \ v^{AY'}_{ZZ'} = -{\sqrt{2}\over\sqrt{\rho^2+X^2}}X^{A}_{\ Z}\delta^{Y'}_{Z'} \cr u^{YY'}_{BZ'}& = {\sqrt{2}\over\sqrt{\rho^2+X^2}}X_{B}^{\ Y}\delta^{Y'}_{Z'} \ \ \ \ \ \ u^{AY'}_{BZ'} = {\rho\over\sqrt{\rho^2+X^2}}\delta^A_{B}\delta^{Y'}_{Z'} \ ,\cr} \eqn\uv $$ and the mass term \Om\ is $$ \Omega ={1\over2}(X^2 + \rho^2) \ . $$ The gauge field $A_{ijAY}$ obtained from \uv\ is simply that of a single instanton on the manifold ${\bf R}^4$ $$ A^{YY'ZZ'}_{AX} = - \epsilon^{Y'Z'} {(\delta^{Z}_{X}X_{A}^{\ \ Y} + \delta^{Y}_{X}X_{A}^{\ \ Z})\over\rho^2+X^2} \ , \eqn\A $$ and the four fermion vertex \Fdef\ is $$ F^{TT'UU'}_{A'YB'Z} = {4\rho^2\over(X^2 + \rho^2)^2} \epsilon_{A'B'} \epsilon^{T'U'} \delta^{T}_{(Y}\delta^{U}_{Z)}\ , \eqn\F $$ which is precisely the field strength of an instanton, justifying our presumptuous notation, and can be easily seen to satisfy \constraint . \subsection{Anomalies} So far we have ignored the possibility of anomalies in the quantum theory. While the original theory \action\ is simply a linear sigma model and therefore possesses no anomalies, this is not the case for the effective theory \effective. It is well known that off-shell $(0,4)$ supersymmetric sigma models suffer from chiral anomalies which break spacetime gauge and coordinate invariance, unless the gauge field can be embedded in the spin connection of the target space. In addition, working in $(0,1)$ superspace only ensures that $(0,1)$ supersymmetry is preserved and there are also extended supersymmetry anomalies as the $(0,4)$ supersymmetry is not preserved. We therefore expect that we will have to add finite local counter terms to \effective\ at all orders of perturbation theory so as to cancel these anomalies. This requires that the spacetime metric and antisymmetric tensor fields become non trivial at higher orders of $\alpha'$, while on the other hand the gauge connection is unaffected [\HPa]. An alternative way of viewing this is to note that although the action \effective\ is classically conformally invariant, when quantized it may not be ultraviolet finite and hence break scale invariance. There is a power counting argument which asserts that off-shell $(0,4)$ supersymmetric models are ultraviolet finite to all orders of perturbation theory [\HPb]. This argument is further complicated by sigma model anomalies and it has been stated that only the non chiral models are ultraviolet finite. Indeed while off-shell $(0,4)$ supersymmetric theories are one loop finite, there is a two loop contribution of the form ${\rm Tr}(R^2 -F^2)$ [\NDL] which certainly does not vanish in general. Normally this leads to non vanishing $\beta$-functions which we must then take into account when determining the conformal fixed point of the renormalization group flow. However, in models with off-shell $(0,4)$ supersymmetry the non vanishing $\beta$-functions can be canceled by redefining the spacetime fields at each order of $\alpha'$, in such a way as to ensure that supersymmetry is preserved in perturbation theory [\HPa]. This has been well studied and verified up to three loops. Thus the ultraviolet divergences which arise in the quantization of off-shell $(0,4)$ models are really an artifact of the use of a renormalization scheme which does not preserve the supersymmetry. The off-shell $(0,4)$ models are ultraviolet finite in an appropriate renormalization scheme. However the model here has only on-shell $(0,4)$ supersymmetry and these finiteness arguments do not immediately apply. At least in the $k=k'=1$ case however the gauge group $SO(4)\cong SU(2)\times SU(2)$ contains a subgroup $Sp(1)\cong SU(2)$ which admits three complex structures obeying the algebra of the quaternions. This endows the target space of the left handed fermions with a hyper Kahler structure and facilitates an off-shell formulation using constrained superfields [\HPc]. We may therefore expect that it is ultraviolet finite in the same manner as the off-shell models described above. In [\HPa] the necessary field redefinitions were derived to order $\alpha'^2$ for $(0,4)$ supersymmetric sigma models. Both the target space metric and antisymmetric tensor field strength receive corrections to all orders in $\alpha'$. Howe and Papadopoulos found that in order to maintain $(0,4)$ supersymmetry in perturbation theory the target space metric (which is flat here at the classical level) must receive corrections in the form of a conformal factor $$ \epsilon_{AB} \epsilon_{YZ} \rightarrow \left(1 - {3\over2}\alpha' f - {3\over16}\alpha'^2 \triangle f + ... \right) \epsilon_{AB} \epsilon_{YZ}\ . \eqn\conf $$ They also showed, up to three loop order, that these redefinitions cancel the ultraviolet divergences which arise when one renormalizes \effective\ using standard $(0,1)$ superspace methods, which do not ensure (0,4) supersymmetry is preserved perturbatively. In addition, the antisymmetric field strength tensor becomes $H = -{3\over4}\alpha'*df$ so as to cancel the gauge anomaly $dH=-{3\alpha'\over4}{\rm Tr} F \wedge F$. Furthermore there are no corrections to the instanton gauge field. For the instanton number one model considered here, Howe and Papadopoulos give the function $f$ as $$ f = -\triangle {\rm ln}(X^2 + \rho^2)\ , $$ where $\triangle$ is the flat space Laplacian. It is a simple matter to calculate the conformal factor \conf\ and hence the target space metric as $$ g_{AYBZ} = \left(1 + 6\alpha'{X^2+2\rho^2 \over(X^2+\rho^2)^2} - 18\alpha'^2 {\rho^4 \over(X^2+\rho^2)^4}+... \right)\epsilon_{AB} \epsilon_{YZ} \ . \eqn\metric $$ To order $\alpha'$ this is the solution of Callan, Harvey and Strominger [\CHS] obtained by solving the first order equations of motion of the 10 dimensional heterotic string (although with $n=6$ rather than $n=8$ in their notation). Thus the target space has been curved around the instanton by stringy effects but remains non singular so long as $\rho \ne 0$. The case $\rho=0$ is of great interest as it provides a string theoretic compactification of instanton moduli space. We will briefly discuss this in the next section. \chapter{Concluding Remarks} In the above we found the order $\alpha'^2$ corrections to the low energy effective action of the ADHM sigma model. Such solutions have been discussed before [\CHS] and we agree with their solution to first order. In our calculations we have expanded in the parameter $$ \alpha'\Omega^{-1}={2\alpha'\over{X^2 +\rho^2}} $$ and hence our approximations are valid for all $X$ if $\rho^2 \gg \alpha'$ and for $X^2 \gg \alpha'$ even when $\rho^2$ is small. An interesting question raised is what are the stringy corrections to the classical instanton in the extreme case that it's size vanishes? One can see from \metric\ that the order $\alpha'$ corrections persist when $\rho=0$ so the effective theory is non trivial. It has been conjectured that there should be a $(4,4)$ supersymmetric sigma model for instantons of zero size [\GS,\GSt] which could be constructed from a massive linear $(4,4)$ supersymmetric model. In [\GSt] the conditions for the ADHM model to possess full (4,4) supersymmetry in the infrared limit were derived. There it was found that the metric must be conformally flat, with the metric satisfying Laplace's equation. This is in agreement with what we have found here in the $\rho=0$ case above (see \conf\ and \metric ) and lends some additional support to this conjecture. If we start with the linear sigma model \action\ with $\rho=0$ the $\lambda_+^{AY}$ fields are massless and decouple from the theory. The vacuum states are defined by $X^{AY}=0$ or $\phi^{A'Y'}=0$ and there is a symmetry between $X^{AY}$ and $\phi^{A'Y'}$. Let us assume we choose the $\phi^{A'Y'}=0$ vacuum. Then as before the fields $X^{AY}$ and $\psi_-^{AY'}$ are massless and the $\phi^{A'Y'}$, $\chi_-^{AY'}$ and $\lambda_+^{YY'}$ fields all have masses $ m\sqrt{X^2/2}$. Upon integrating out the massive fields we would simply obtain a free field theory, which trivially possesses $(4,4)$ supersymmetry. At the degenerate vacuum $X^{AY}=0$ however, all fields are massless and there is a single interaction term $m\lambda_{+YY'}\phi^{\ \ Y'}_{A'}\psi_-^{A'Y}$. Thus the moduli space of vacua does not have a manifold structure. For $X^{AY} \ne 0$ the vacuum states are simply ${\bf R}^4$ but at the point $X^{AY} = 0$ lies another entire copy of ${\bf R}^4$ (parameterized by the $\phi^{A'Y'}$). This odd state of affairs is smoothly resolved if we first construct the effective theory and then take the limit of vanishing instanton size. Let us now take the limit $\rho \rightarrow 0$ of the effective action \effective . It should be noted that the Yang-Mills instanton has shrunk to zero size but it has not disappeared in the sense that the topological charge remains equal to one. Unfortunately our expressions are not a priori valid near $X=0$. Nevertheless we will try to shed some light about what the complete string theory solution could be in that region. When $\rho$ vanishes both the field strength \Fdef\ and the $O(\alpha')$ sigma model anomaly vanish. We are however, still left with a non trivial metric and anti symmetric tensor. It seems reasonable to assume then that all the anomalies are canceled by these. The metric then has the exact conformal factor $$ g_{\mu\nu} = \left(1 - {3\alpha'\over2}f\right)\delta_{\mu\nu} \ , \eqn\cfact $$ and antisymmetric field $$ H_{\mu\nu\rho} = -{3\alpha'\over4}\epsilon_{\mu\nu\rho\lambda}\partial^{\lambda}f \ , \eqn\H $$ where $f = -4/X^2$ and we have switched to a more convenient notation. This geometry is similar to the one discussed by Callan, Harvey and Strominger [\CHS], although the anti symmetric field is not the same and leads to a different interpretation in the limit $X^2 \rightarrow 0$ as we will shortly see. The target space is non singular, asymptotically flat and has a semi-infinite tube with asymptotic radius $\sqrt{6\alpha'}$, centered around the instanton. That is to say the apparent singularity at $X^2=0$ in \metric\ is pushed off to an "internal infinity" down the infinite tube. Thus the problematic $X^{AY}=0$ vacua are pushed an infinite distance away and the manifold structure is preserved. The resolution of this description with the non manifold picture described above has been discussed by Witten [\Wittenb]. In the limit $X^2 \ll \alpha'$ the modified spin connection with torsion becomes, where $\alpha,\beta$ are vierbein indices, $$ \eqalign{ \omega_{\mu}^{(-)\alpha\beta} &\equiv \omega_{\mu}^{\alpha\beta} + H_{\mu}^{\ \alpha\beta} \cr &= -(\delta^\alpha_{\mu} \delta^\beta_{\nu} -\delta^\alpha_{\nu} \delta^\beta_{\mu} + \epsilon_{\mu\nu}^{\ \ \ \alpha\beta}){X^{\nu}\over X^2} \cr } \ , \eqn\ominus $$ which is a flat connection! That is to say far down the infinite tube the torsion parallelizes the manifold (which is asymptotically $S^3 \times {\bf R}$ and is indeed parallelizable). The gauge connection is also flat (for $X^{AY} \ne 0$) and can therefore be embedded into the generalized spin connection \ominus\ (they are both $so(4)$ valued). The low energy effective theory therefore possesses $(4,4)$ supersymmetry in the region $X^2 \rightarrow 0$ and is free of anomalies there. This supports our assumption that the anomalies are canceled and the expressions \cfact\ and \H\ are exact, at least in this region. For the region $X^2 \gg \alpha'$ our perturbative expansion is valid and the the theory only possesses $(0,4)$ supersymmetry since the gauge connection can not be embedded into the spin connection. Although in a similar spirit in the limit $X^2 \rightarrow \infty$ the curvatures vanish and the theory is free and again has $(4,4)$ supersymmetry. In a sense then the $\rho=0$ ADHM instanton can be viewed as a soliton in the space of string vacua interpolating between two $(4,4)$ supersymmetric sigma models, just as the target space can be viewed as interpolating between two supersymmetric ground states of supergravity [\GT]. However, since we still have instanton number one, the vector bundle is non trivial whereas the tangent bundle is trivial. Thus in the region $X^2 \rightarrow 0$ we can only identify the spin connection with the gauge connection locally. $(4,4)$ supersymmetry is then broken by global, non perturbative effects back to $(0,4)$ supersymmetry. This is reflected by the observation [\GT] that the $S^3 \times M^7$ compactification of $D=10$ supergravity breaks half the supersymmetry. I would like to thank G. Papadopoulos and P.K. Townsend for their advice and Trinity College Cambridge for financial support. \refout \end . Therefore, in the region $X^2 \ll \alpha'$, the torsion parallelizes the manifold
\section{Introduction} This talk is organized as follows. First, we briefly discuss the supersymmetric standard model, in order to introduce the parameter space we are considering. Next, we discuss the prediction for the strong coupling constant and the bottom pole mass with and without including GUT threshold effects, in the small $\tan\beta$ case. Lastly we discuss our results in the large $\tan\beta$ regime. The minimal supersymmetric model is attractive in that it solves the hierarchy problem. It does this, however, at the expense of doubling the number of degrees of freedom of the standard model. This is problematic, as supersymmetry must be broken and hence the number of new parameters needed to describe the model is in general quite large. An organizing principle is needed in order to reduce the number of parameters. In a minimal supergravity scenario the number of supersymmetry breaking parameters needed to describe the supersymmetric model are few: a universal scalar mass $M_0$, a universal gaugino mass $M_{1/2}$, a universal trilinear scalar coupling $A_0$, and a bilinear scalar coupling $B$. In addition there is a supersymmetric Higgs mass term, $\mu$. Given values for these five parameters at the GUT scale, we use the renormalization group equations~\cite{RGEs} (RGE's) to determine the various particle masses and couplings at the weak scale. For a large top-quark mass, one of the Higgs boson masses is driven negative, and the radiative breaking of $SU(2)\times U(1)$ symmetry becomes manifest. The Higgs bosons obtain vev's $v_1$ and $v_2$, and we can determine the $Z$-boson mass. In practice it is convenient to assume electroweak symmetry breaks radiatively and to take $M_Z$ as an input parameter, as well as the ratio of vev's $\tan\beta\equiv v_2/v_1$. We then determine $\mu^2$ and $B$ from the symmetry breaking conditions. To summarize, then, the supersymmetric model is parametrized by $$M_0,\ \ M_{1/2},\ \ A_0,\ \ \tan\beta,\ \ {\rm sign}(\mu).$$ The exact one-loop corrections to the masses, gauge couplings and Yukawa couplings of the minimal supersymmetric model are described in Ref. [2]. These corrections are essential ingredients for accurate tests of grand unification. They allow one to extract the underlying \mbox{\footnotesize$\overline{\it DR}$}~ parameters from a given set of measured observables. The \mbox{\footnotesize$\overline{\it DR}$}~ parameters can then be run up to a high scale to explore the consequences of different unification hypotheses. Alternatively, the radiative corrections can be used to translate various limits into excluded regions of the \mbox{\footnotesize$\overline{\it DR}$}~ parameter space. This is illustrated in Fig.~1, where we show the excluded region of the $M_0, \ M_{1/2}$ parameter space at the one-loop level, from current experimental constraints. \begin{figure}[t] \epsfysize=3in \epsffile[30 495 600 755]{f1.ps} \begin{center} \parbox{5.5in}{ \caption[]{\footnotesize Excluded region (shaded) of the $M_0,\ M_{1/2}$ plane, for $\tan\beta=2,\ m_t=175$ GeV, $A_0=0$, and $\alpha_s=0.117$. All masses are evaluated at one-loop. The symbols indicate which experimental constraint is relevant: $\chi^+\Rightarrow m_{\chi^+} > 47$ GeV; $\tilde g\Rightarrow m_{\tilde g}>125$ GeV; $\tilde\nu\Rightarrow m_{\tilde\nu}>42$ GeV; $h\Rightarrow m_h>60$ GeV.}} \end{center} \end{figure} In the following, we treat all supersymmetric threshold corrections in a complete one-loop analysis.\footnote{See Chankowski et al.~\cite{Chankowski} for a similar treatment of finite corrections to $\sin^2\theta_W$.} Our work stands in contrast to most previous studies, which are based on the ``leading logarithm approximation." For the gauge coupling threshold corrections, this approximation involves taking the standard-model value of $\sin^2\theta_W$ and adding the logarithmic parts of the SUSY threshold corrections. The approximation works well if all of the SUSY particle masses are much greater than $M_Z$, in which case the decoupling theorem implies that the finite effects of the SUSY particles are negligible for all low-energy observables. However, in realistic models it is not unusual for the supersymmetric spectrum to contain light particles of order the $Z$-mass. In this case the leading logarithm approximation breaks down. This is illustrated in Fig.~2, where we compare the value of $\alpha_s$ and $m_b$ in the leading logarithm approximation (LLA) with the value obtained in the full calculation. In this talk we use the full set of one-loop radiative corrections to evaluate the \mbox{\footnotesize$\overline{\it DR}$}~ gauge and Yukawa couplings. The \mbox{\footnotesize$\overline{\it DR}$}~ couplings serve as the boundary conditions for the two-loop gauge and Yukawa coupling renormalization group equations, which determine the couplings at very high scales. \begin{figure}[t] \epsfysize=3in \epsffile[35 475 600 735]{f2.ps} \begin{center} \parbox{5.5in}{ \caption[]{\footnotesize Comparison of (a) $\alpha_s$ and (b) $m_b$ in the leading logarithm approximation (LLA) versus the full one-loop calculation, for $M_0=60$ GeV, $\tan\beta=2, \ m_t=175$ GeV, $A_0=0$, and $\mu>0$. The dashed line shows the result if the non-universal corrections are neglected.}} \end{center} \end{figure} In what follows, we have converted the strong coupling to the \mbox{\footnotesize$\overline{\it MS}$}~ scheme so that by $\alpha_s$ we refer to the standard \mbox{\footnotesize$\overline{\it MS}$}~ value evaluated at the scale $M_Z$. By $m_b$ we refer to the bottom-quark pole mass. \section{Small $\tan\beta$} As a reference point, we show in Fig.~3 contours of $m_b$ and $\alpha_s$ in the $M_0,\ M_{1/2}$ plane, with no GUT thresholds, $\tan\beta=2$, $m_t=175$ GeV, and $A_0$=0. We confine our attention to the region of the theory which is more natural, i.e. to the region where the superpartner masses are less than about 1 TeV. We find the strong coupling is large ($\alpha_s>0.127$) compared to the PDG value \cite{pdg} $\alpha_s=0.117\pm0.005$. Similarly, the bottom mass is quite large ($m_b>6$ GeV), far outside the preferred region which we take to be $4.7 < m_b < 5.2$ GeV. \begin{figure}[t] \epsfysize=3in \epsffile[30 468 600 728]{f3.ps} \begin{center} \parbox{5.5in}{ \caption[]{\footnotesize Contours of $\alpha_s$ (solid lines) and $m_b$ (dashed) in the $M_0$, $M_{1/2}$ plane, with $\tan\beta=2$, $m_t=175$ GeV, and $A_0=0$. The upper left hand corners are excluded on cosmological grounds (charged LSP) and the lower shaded regions are excluded by particle searches. The $m_b$ contours are labeled in GeV.}} \end{center} \end{figure} The experimental uncertainty in the determination of $\alpha_s$ is primarily due to the uncertainty in determining the electromagnetic coupling at the $Z$-scale. We use the value recently determined by Eidelman and Jegerlehner \cite{Eidelman}. The one-sigma uncertainty in our input $\alpha_{EM}(M_Z)$ results in a one-sigma uncertainty in our output $\alpha_s$ of about $\pm0.001$, and an uncertainty of typically 0.07 GeV in our bottom mass prediction. Martin and Zeppenfeld \cite{Martin} and Swartz \cite{Swartz} have also performed analyses to determine $\alpha_{EM}(M_Z)$. The central value for $\alpha_s$ increases by about 0.001 if we use the value of $\alpha_{\rm EM}(M_Z)$ as determined by Martin and Zeppenfeld, and it increases by about 0.002 if we use the value of $\alpha_{EM}(M_Z)$ from Swartz. The bottom mass is reduced if we go further into the small $\tan\beta$ region, due to the large top-quark Yukawa coupling in the bottom mass renormalization group equation. We consider the case where the top Yukawa is as large as possible; we set $\lambda_t(M_{\rm GUT})=3$, which is on the verge of the nonperturbative regime. In this case we will obtain the smallest possible bottom mass. As seen in Fig.~4, for $m_t=180$ GeV the bottom mass is less than 5.2 GeV and the strong coupling is less than 0.127 only if the squark masses are in the TeV region. \begin{figure}[t] \epsfysize=3.5in \epsffile[25 398 600 718]{f4.ps} \begin{center} \parbox{5.5in}{ \caption[]{\footnotesize The bottom-quark mass and $\alpha_s$ vs. $m_t$ for the case of no GUT-scale thresholds, for various values of $\tan\beta$, with $A_0=0$, $\mu>0$, and $\lambda_t(M_{\rm GUT})=3$. The circles on the right (solid) leg in each pair of lines corresponds to $M_{1/2}$ equal to (from top to bottom) 60, 100, 200, 500 GeV, with $M_0$ fixed at 100 GeV. The circles on the left (dashed) leg corresponds to $M_0$ equal to 100, 200, 400 and 1000 GeV, with $M_{1/2}=100$ GeV. The horizontal dot-dashed lines indicate $m_b=5.2$ GeV and $\alpha_s=0.127$. The $\times$'s mark points with one-loop Higgs mass $m_h<60$ GeV.}} \end{center} \end{figure} If we consider particular GUT models, we can determine whether the GUT threshold corrections to the gauge and Yukawa couplings can help improve the situation. We parametrize the GUT threshold corrections by $\varepsilon_g$ and $\varepsilon_b$, where $$ g_3(M_{\rm GUT}) = g_{\rm GUT}(M_{\rm GUT})\left( 1+\varepsilon_g\right)\ ,$$ $$ \lambda_b(M_{\rm GUT})=\lambda_\tau(M_{\rm GUT})\left( 1+\varepsilon_b\right)\ ,$$ where $\lambda_b$ and $\lambda_\tau$ are the $b$- and $\tau$-Yukawa couplings, and $M_{\rm GUT}$ is defined as the scale at which $g_1$ and $g_2$ meet, $g_{\rm GUT}\equiv g_1(M_{\rm GUT})=g_2(M_{\rm GUT})$. A smaller value of $\alpha_s$ requires $\varepsilon_g<0$. In the small $\tan\beta$ region the bottom mass is tightly correlated with the value of the strong coupling. Hence, setting $\varepsilon_g<0$ reduces both $\alpha_s$ and $m_b$. In fact, the bottom mass is an order of magnitude more sensitive to $\varepsilon_g$ than to $\varepsilon_b$. In what follows we examine the GUT corrections in two SU(5) GUT models. \begin{figure}[t] \epsfysize=3in \epsffile[30 468 600 728]{f5.ps} \begin{center} \parbox{5.5in}{ \caption[]{\footnotesize Contours of the smallest possible $\alpha_s$ (solid lines) and $m_b$ (dashed) consistent with nucleon decay in (a) minimal SU(5), and (b) missing doublet SU(5), with $m_t=175$ GeV, $A_0=0$ and $\mu>0$. In (a) $\lambda_t(M_{\rm GUT})=3$ ($\tan\beta\simeq1.4$) and in (b) $\tan\beta=2$. The shaded regions are phenomenologically excluded. The $m_b$ contours are labeled in GeV.}} \end{center} \end{figure} \begin{figure}[t] \epsfysize=3in \epsffile[30 468 600 728]{f6.ps} \begin{center} \parbox{5.5in}{ \caption[]{\footnotesize Contours of $\alpha_s$ (solid) and $m_b$ (dashed, labeled in GeV) with no GUT thresholds, $m_t=175$ GeV, $A_0=0$ and $\tan\beta=30$.}} \end{center} \end{figure} \begin{figure}[t] \epsfysize=3in \epsffile[30 468 600 728]{f7.ps} \begin{center} \parbox{5.5in}{ \caption[]{\footnotesize Contours of $\alpha_s$ (solid) and $m_b$ (dashed, labeled in GeV) in missing doublet SU(5), with $m_t=175$ GeV, $A_0=0$ and $\tan\beta=30$.}} \end{center} \end{figure} In the minimal SU(5) model \cite{minSUfive}, the gauge coupling threshold correction $\varepsilon_g'$ is given by \cite{minSUfive epsg} \begin{equation} \varepsilon'_g\ =\ {3g_{\rm GUT}^2\o40\pi^2}\,\log\left(M_{H_3}\over M_{\rm GUT}\right)\ , \label{epsprime} \end{equation} where $M_{H_3}$ is the mass of the color-triplet Higgs particle that mediates nucleon decay. From this expression, we see that $\varepsilon'_g < 0$ whenever $M_{H_3} < M_{\rm GUT}$. However, $M_{H_3}$ is bounded from below by proton decay experiments. The $M_{H_3}$ mass limit is of the form \cite{Hisano} $$ M_{H_3} > {\cal M}\ {|1+y^{tK}|\over\sin2\beta} f(\tilde w, \tilde d, \tilde u,\tilde e) $$ where ${\cal M}$ is a nuclear matrix element, $y^{tK}$ parametrizes the amount of third generation mixing, and $f$ is a function of the wino, squark and slepton masses. For the conservative choices ${\cal M}=0.003$ GeV$^3$ and $|1+y^{tK}|=0.4$ we find that $M_{H_3}^{\rm min}>M_{\rm GUT}$ unless $M_0>500$ GeV and $M_{1/2} \ll M_0$. Thus, in most of the parameter space, $\varepsilon_g'>0$. For this reason, in minimal SU(5), $\alpha_s$ is typically even larger than in the case of no GUT thresholds, as illustrated in Fig.~5(a). In order to obtain the smallest possible $\alpha_s$ and $m_b$, we have set $\lambda_t(M_{\rm GUT})=3$ in Fig.~5(a). Thus we end up with rather small values for $\tan\beta$ ($\sim 1.3$-1.6), and the Higgs mass constraint rules out a large part of parameter space. Only in the region $M_0 \gg M_{1/2}$, where the proton decay amplitude is suppressed, is the strong coupling reduced relative to the case with no GUT thresholds. The smallest value of $\alpha_s$ occurs in this region, a somewhat large value of 0.124. The bottom-quark mass is similarly on the high side of the preferred region. We have applied the most favorable Yukawa correction $\varepsilon_b$ given in Wright~\cite{wright}, subject to Yukawa coupling perturbativity constraints (see Bagger et al.~\cite{bmp}). The missing-doublet model is an alternative SU(5) theory in which the heavy color-triplet Higgs particles are split naturally from the light Higgs doublets \cite{missing-doublet}. In this model the GUT gauge threshold correction is given by \cite{Yamada} \begin{equation} \varepsilon_g^{\prime\prime}\ = \ {3g_{\rm GUT}^2\o40\pi^2}\,\Biggl\{\log\left(M_{H_3}^{\rm eff} \over M_{\rm GUT}\right) - {25\o2}\log5 + 15\log2\Biggr\}\ \simeq \ \varepsilon'_g - 4\%\ . \label{mdmodel} \end{equation} Thus, for fixed $M_{H_3}$, the missing-doublet model has the same threshold correction as the minimal SU(5) model, minus 4\%. In eq.~(\ref{mdmodel}), $M_{H_3}^{\rm eff}$ is the effective mass that enters into the proton decay amplitude, so the bounds on $M_{H_3}$ in the minimal SU(5) model also apply to $M_{H_3}^{\rm eff}$ in the missing-doublet model. The large negative correction in eq.~(\ref{mdmodel}) is due to the mass splitting in the {\bf 75} representation, and gives rise to much smaller values for $\alpha_s$ and $m_b$. This is illustrated in Fig.~5(b), where we show contours of $\alpha_s$ and $m_b$ in the $M_0,\ M_{1/2}$ plane, with $M_{H_3}^{\rm eff}=M_{H_3}^{\rm min}$, at $\tan\beta=2$. We find (even without going into the far infrared top Yukawa fixed point region) values of both the strong coupling and the bottom mass near their central values. \section{Large $\tan\beta$} Again, as a reference point, we show in Fig.~6 the strong coupling and bottom mass prediction with no GUT corrections, for $A_0=0,\ m_t=175$ GeV, at $\tan\beta=30$. The $\alpha_s$ predictions are not significantly different from the small $\tan\beta$ case. The bottom mass is significantly influenced by the large finite corrections~\cite{mb}, $${\Delta m_b\over m_b}\sim -{\mu\tan\beta\over16\pi^2m_{\tilde q}^2} \left({8\over3}g_3^2 m_{\tilde g} + \lambda_t^2 A_t\right)\ ,$$ so much so, that for the case $\mu<0$ the bottom mass is hopelessly large. However, these help to reduce $m_b$ when $\mu>0$. Even so, we see in Fig. 6(a) that $m_b > 5.5$ GeV. GUT threshold corrections are needed to reduce both $\alpha_s$ and $m_b$ further. In minimal SU(5) the proton decay rate is enhanced at large $\tan\beta$, so the triplet Higgs mass $M_{H_3}$ is forced to be quite large. Hence, we have a large and positive GUT correction $\varepsilon_g'$ (Eq. (\ref{epsprime})) which forces the strong coupling to unacceptably large values (\roughly{>}0.14). Hence we conclude that minimal SU(5) is ruled out at large $\tan\beta$. In the missing doublet model the large triplet Higgs mass correction is adequately compensated by the constant $-4\%$ correction of Eq. (\ref{mdmodel}). As shown in Fig. 7, this cancellation results in near central values for $\alpha_s$ and $m_b$. However, $\mu>0$ is clearly required when $\tan\beta$ becomes large, as the large finite corrections have the wrong sign in the $\mu<0$ case, yielding values of $m_b$ which are unacceptably large. \section{Conclusion} \vspace{-.25cm} In this talk we have presented results from a complete calculation of the one-loop corrections to the masses, gauge, and Yukawa couplings in the MSSM. We have seen that such a calculation allows us to reliably investigate various unified models to determine whether they are compatible with current experimental data. In particular, we found that the finite SUSY corrections, which are neglected in the leading logarithm approximation, can substantially increase the prediction for $\alpha_s$ and $m_b$ when some of the SUSY partner masses are lighter than or of order $M_Z$. For small $\tan\beta$, we found that in the minimal SU(5) model, $\alpha_s$ was somewhat large ($\alpha_s>0.124$ with $m_{\tilde q}<1$ TeV) and $m_b$ was larger than 5 GeV. The missing doublet model gave much lower values of $\alpha_s$ and $m_b$, near their central values. For large $\tan\beta$, the minimal SU(5) model GUT threshold correction became quite large and positive, and this resulted in unacceptably large values of the strong coupling ($\alpha_s\roughly{>}0.14$). The missing doublet model, on the other hand, had no trouble in accomodating the central values of the strong coupling and $m_b$. The values of $m_b$ were largely determined by important finite corrections, and we required $\mu>0$ in order for these corrections to result in acceptably small values of the bottom-quark mass. \section{Acknowledgements} I would like to thank my collaborators Jonathan Bagger and Konstantin Matchev. This work was supported by the U.S. National Science Foundation under grant NSF-PHY-9404057. \section{References}